fbpx
Wikipedia

Acid dissociation constant

In chemistry, an acid dissociation constant (also known as acidity constant, or acid-ionization constant; denoted ) is a quantitative measure of the strength of an acid in solution. It is the equilibrium constant for a chemical reaction

known as dissociation in the context of acid–base reactions. The chemical species HA is an acid that dissociates into A, the conjugate base of the acid and a hydrogen ion, H+.[a] The system is said to be in equilibrium when the concentrations of its components will not change over time, because both forward and backward reactions are occurring at the same rate.[1]

The dissociation constant is defined by[b]

or

where quantities in square brackets represent the concentrations of the species at equilibrium.[c][2] As a simple example for a weak acid with Ka = 10−5, log Ka is the exponent which is -5, so that pKa = 5. And for acetic acid with Ka = 1.8 x 10−5, pKa is close to 5. A higher Ka corresponds to a stronger acid which is more dissociated at equilibrium. For the more convenient logarithmic scale, a lower pKa means a stronger acid.

Theoretical background edit

The acid dissociation constant for an acid is a direct consequence of the underlying thermodynamics of the dissociation reaction; the pKa value is directly proportional to the standard Gibbs free energy change for the reaction. The value of the pKa changes with temperature and can be understood qualitatively based on Le Châtelier's principle: when the reaction is endothermic, Ka increases and pKa decreases with increasing temperature; the opposite is true for exothermic reactions.

The value of pKa also depends on molecular structure of the acid in many ways. For example, Pauling proposed two rules: one for successive pKa of polyprotic acids (see Polyprotic acids below), and one to estimate the pKa of oxyacids based on the number of =O and −OH groups (see Factors that affect pKa values below). Other structural factors that influence the magnitude of the acid dissociation constant include inductive effects, mesomeric effects, and hydrogen bonding. Hammett type equations have frequently been applied to the estimation of pKa.[3][4]

The quantitative behaviour of acids and bases in solution can be understood only if their pKa values are known. In particular, the pH of a solution can be predicted when the analytical concentration and pKa values of all acids and bases are known; conversely, it is possible to calculate the equilibrium concentration of the acids and bases in solution when the pH is known. These calculations find application in many different areas of chemistry, biology, medicine, and geology. For example, many compounds used for medication are weak acids or bases, and a knowledge of the pKa values, together with the octanol-water partition coefficient, can be used for estimating the extent to which the compound enters the blood stream. Acid dissociation constants are also essential in aquatic chemistry and chemical oceanography, where the acidity of water plays a fundamental role. In living organisms, acid–base homeostasis and enzyme kinetics are dependent on the pKa values of the many acids and bases present in the cell and in the body. In chemistry, a knowledge of pKa values is necessary for the preparation of buffer solutions and is also a prerequisite for a quantitative understanding of the interaction between acids or bases and metal ions to form complexes. Experimentally, pKa values can be determined by potentiometric (pH) titration, but for values of pKa less than about 2 or more than about 11, spectrophotometric or NMR measurements may be required due to practical difficulties with pH measurements.

Definitions edit

According to Arrhenius's original molecular definition, an acid is a substance that dissociates in aqueous solution, releasing the hydrogen ion H+ (a proton):[5]

 

The equilibrium constant for this dissociation reaction is known as a dissociation constant. The liberated proton combines with a water molecule to give a hydronium (or oxonium) ion H3O+ (naked protons do not exist in solution), and so Arrhenius later proposed that the dissociation should be written as an acid–base reaction:

 
 
Acetic acid, a weak acid, donates a proton (hydrogen ion, highlighted in green) to water in an equilibrium reaction to give the acetate ion and the hydronium ion. Red: oxygen, black: carbon, white: hydrogen.

Brønsted and Lowry generalised this further to a proton exchange reaction:[6][7][8]

 

The acid loses a proton, leaving a conjugate base; the proton is transferred to the base, creating a conjugate acid. For aqueous solutions of an acid HA, the base is water; the conjugate base is A and the conjugate acid is the hydronium ion. The Brønsted–Lowry definition applies to other solvents, such as dimethyl sulfoxide: the solvent S acts as a base, accepting a proton and forming the conjugate acid SH+.

 

In solution chemistry, it is common to use H+ as an abbreviation for the solvated hydrogen ion, regardless of the solvent. In aqueous solution H+ denotes a solvated hydronium ion rather than a proton.[9][10]

The designation of an acid or base as "conjugate" depends on the context. The conjugate acid BH+ of a base B dissociates according to

 

which is the reverse of the equilibrium

 

The hydroxide ion OH, a well known base, is here acting as the conjugate base of the acid water. Acids and bases are thus regarded simply as donors and acceptors of protons respectively.

A broader definition of acid dissociation includes hydrolysis, in which protons are produced by the splitting of water molecules. For example, boric acid (B(OH)3) produces H3O+ as if it were a proton donor,[11] but it has been confirmed by Raman spectroscopy that this is due to the hydrolysis equilibrium:[12]

 

Similarly, metal ion hydrolysis causes ions such as [Al(H2O)6]3+ to behave as weak acids:[13]

 

According to Lewis's original definition, an acid is a substance that accepts an electron pair to form a coordinate covalent bond.[14]

Equilibrium constant edit

An acid dissociation constant is a particular example of an equilibrium constant. The dissociation of a monoprotic acid, HA, in dilute solution can be written as

 

The thermodynamic equilibrium constant   can be defined by[15]

 

where {X} represents the activity, at equilibrium, of the chemical species X.   is dimensionless since activity is dimensionless. Activities of the products of dissociation are placed in the numerator, activities of the reactants are placed in the denominator. See activity coefficient for a derivation of this expression.

 
Variation of pKa of acetic acid with ionic strength.

Since activity is the product of concentration and activity coefficient (γ) the definition could also be written as

 

where   represents the concentration of HA and   is a quotient of activity coefficients.

To avoid the complications involved in using activities, dissociation constants are determined, where possible, in a medium of high ionic strength, that is, under conditions in which   can be assumed to be always constant.[15] For example, the medium might be a solution of 0.1 molar (M) sodium nitrate or 3 M potassium perchlorate. With this assumption,

 
 

is obtained. Note, however, that all published dissociation constant values refer to the specific ionic medium used in their determination and that different values are obtained with different conditions, as shown for acetic acid in the illustration above. When published constants refer to an ionic strength other than the one required for a particular application, they may be adjusted by means of specific ion theory (SIT) and other theories.[16]

Cumulative and stepwise constants edit

A cumulative equilibrium constant, denoted by   is related to the product of stepwise constants, denoted by   For a dibasic acid the relationship between stepwise and overall constants is as follows

 
 
 

Note that in the context of metal-ligand complex formation, the equilibrium constants for the formation of metal complexes are usually defined as association constants. In that case, the equilibrium constants for ligand protonation are also defined as association constants. The numbering of association constants is the reverse of the numbering of dissociation constants; in this example  

Association and dissociation constants edit

When discussing the properties of acids it is usual to specify equilibrium constants as acid dissociation constants, denoted by Ka, with numerical values given the symbol pKa.

 

On the other hand, association constants are used for bases.

 

However, general purpose computer programs that are used to derive equilibrium constant values from experimental data use association constants for both acids and bases. Because stability constants for a metal-ligand complex are always specified as association constants, ligand protonation must also be specified as an association reaction.[17] The definitions show that the value of an acid dissociation constant is the reciprocal of the value of the corresponding association constant:

 
 
 

Notes

  1. For a given acid or base in water, pKa + pKb = pKw, the self-ionization constant of water.
  2. The association constant for the formation of a supramolecular complex may be denoted as Ka; in such cases "a" stands for "association", not "acid".
  3. For polyprotic acids, the numbering of stepwise association constants is the reverse of the numbering of the dissociation constants. For example, for phosphoric acid (details in the polyprotic acids section below):
 

Temperature dependence edit

All equilibrium constants vary with temperature according to the van 't Hoff equation[18]

 

  is the gas constant and   is the absolute temperature. Thus, for exothermic reactions, the standard enthalpy change,  , is negative and K decreases with temperature. For endothermic reactions,   is positive and K increases with temperature.

The standard enthalpy change for a reaction is itself a function of temperature, according to Kirchhoff's law of thermochemistry:

 

where   is the heat capacity change at constant pressure. In practice   may be taken to be constant over a small temperature range.

Dimensionality edit

In the equation

 

Ka appears to have dimensions of concentration. However, since  , the equilibrium constant,  , cannot have a physical dimension. This apparent paradox can be resolved in various ways.

  1. Assume that the quotient of activity coefficients has a numerical value of 1, so that   has the same numerical value as the thermodynamic equilibrium constant  .
  2. Express each concentration value as the ratio c/c0, where c0 is the concentration in a [hypothetical] standard state, with a numerical value of 1, by definition.[19]
  3. Express the concentrations on the mole fraction scale. Since mole fraction has no dimension, the quotient of concentrations will, by definition, be a pure number.

The procedures, (1) and (2), give identical numerical values for an equilibrium constant. Furthermore, since a concentration   is simply proportional to mole fraction   and density  :

 

and since the molar mass   is a constant in dilute solutions, an equilibrium constant value determined using (3) will be simply proportional to the values obtained with (1) and (2).

It is common practice in biochemistry to quote a value with a dimension as, for example, "Ka = 30 mM" in order to indicate the scale, millimolar (mM) or micromolar (μM) of the concentration values used for its calculation.

Strong acids and bases edit

An acid is classified as "strong" when the concentration of its undissociated species is too low to be measured.[6] Any aqueous acid with a pKa value of less than 0 is almost completely deprotonated and is considered a strong acid.[20] All such acids transfer their protons to water and form the solvent cation species (H3O+ in aqueous solution) so that they all have essentially the same acidity, a phenomenon known as solvent leveling.[21][22] They are said to be fully dissociated in aqueous solution because the amount of undissociated acid, in equilibrium with the dissociation products, is below the detection limit. Likewise, any aqueous base with an association constant pKb less than about 0, corresponding to pKa greater than about 14, is leveled to OH and is considered a strong base.[22]

Nitric acid, with a pK value of around −1.7, behaves as a strong acid in aqueous solutions with a pH greater than 1.[23] At lower pH values it behaves as a weak acid.

pKa values for strong acids have been estimated by theoretical means.[24] For example, the pKa value of aqueous HCl has been estimated as −9.3.

Monoprotic acids edit

 
Variation of the % formation of a monoprotic acid, AH, and its conjugate base, A, with the difference between the pH and the pKa of the acid.

After rearranging the expression defining Ka, and putting pH = −log10[H+], one obtains[25]

 

This is the Henderson–Hasselbalch equation, from which the following conclusions can be drawn.

  • At half-neutralization the ratio [A]/[HA] = 1; since log(1) = 0, the pH at half-neutralization is numerically equal to pKa. Conversely, when pH = pKa, the concentration of HA is equal to the concentration of A.
  • The buffer region extends over the approximate range pKa ± 2. Buffering is weak outside the range pKa ± 1. At pH ≤ pKa − 2 the substance is said to be fully protonated and at pH ≥ pKa + 2 it is fully dissociated (deprotonated).
  • If the pH is known, the ratio may be calculated. This ratio is independent of the analytical concentration of the acid.

In water, measurable pKa values range from about −2 for a strong acid to about 12 for a very weak acid (or strong base).

A buffer solution of a desired pH can be prepared as a mixture of a weak acid and its conjugate base. In practice, the mixture can be created by dissolving the acid in water, and adding the requisite amount of strong acid or base. When the pKa and analytical concentration of the acid are known, the extent of dissociation and pH of a solution of a monoprotic acid can be easily calculated using an ICE table.

Polyprotic acids edit

 
Phosphoric acid speciation

A polyprotic acid is a compound which may lose more than 1 proton. Stepwise dissociation constants are each defined for the loss of a single proton. The constant for dissociation of the first proton may be denoted as Ka1 and the constants for dissociation of successive protons as Ka2, etc. Phosphoric acid, H3PO4, is an example of a polyprotic acid as it can lose three protons.

Equilibrium pK definition and value[26]
   
   
   

When the difference between successive pK values is about four or more, as in this example, each species may be considered as an acid in its own right;[27] In fact salts of H
2
PO
4
may be crystallised from solution by adjustment of pH to about 5.5 and salts of HPO2−4 may be crystallised from solution by adjustment of pH to about 10. The species distribution diagram shows that the concentrations of the two ions are maximum at pH 5.5 and 10.

 
% species formation calculated with the program HySS for a 10 millimolar solution of citric acid. pKa1 = 3.13, pKa2 = 4.76, pKa3 = 6.40.

When the difference between successive pK values is less than about four there is overlap between the pH range of existence of the species in equilibrium. The smaller the difference, the more the overlap. The case of citric acid is shown at the right; solutions of citric acid are buffered over the whole range of pH 2.5 to 7.5.

According to Pauling's first rule, successive pK values of a given acid increase (pKa2 > pKa1).[28] For oxyacids with more than one ionizable hydrogen on the same atom, the pKa values often increase by about 5 units for each proton removed,[29][30] as in the example of phosphoric acid above.

It can be seen in the table above that the second proton is removed from a negatively charged species. Since the proton carries a positive charge extra work is needed to remove it, which is why pKa2 is greater than pKa1. pKa3 is greater than pKa2 because there is further charge separation. When an exception to Pauling's rule is found, it indicates that a major change in structure is also occurring. In the case of VO+2(aq), the vanadium is octahedral, 6-coordinate, whereas vanadic acid is tetrahedral, 4-coordinate. This means that four "particles" are released with the first dissociation, but only two "particles" are released with the other dissociations, resulting in a much greater entropy contribution to the standard Gibbs free energy change for the first reaction than for the others.

Equilibrium pKa
   
   
   
   

Isoelectric point edit

For substances in solution, the isoelectric point (pI) is defined as the pH at which the sum, weighted by charge value, of concentrations of positively charged species is equal to the weighted sum of concentrations of negatively charged species. In the case that there is one species of each type, the isoelectric point can be obtained directly from the pK values. Take the example of glycine, defined as AH. There are two dissociation equilibria to consider.

 
 

Substitute the expression for [AH] from the second equation into the first equation

 

At the isoelectric point the concentration of the positively charged species, AH+2, is equal to the concentration of the negatively charged species, A, so

 

Therefore, taking cologarithms, the pH is given by

 

pI values for amino acids are listed at proteinogenic amino acid. When more than two charged species are in equilibrium with each other a full speciation calculation may be needed.

Bases and basicity edit

The equilibrium constant Kb for a base is usually defined as the association constant for protonation of the base, B, to form the conjugate acid, HB+.

 

Using similar reasoning to that used before

 

Kb is related to Ka for the conjugate acid. In water, the concentration of the hydroxide ion, [OH], is related to the concentration of the hydrogen ion by Kw = [H+][OH], therefore

 

Substitution of the expression for [OH] into the expression for Kb gives

 

When Ka, Kb and Kw are determined under the same conditions of temperature and ionic strength, it follows, taking cologarithms, that pKb = pKw − pKa. In aqueous solutions at 25 °C, pKw is 13.9965,[31] so

 

with sufficient accuracy for most practical purposes. In effect there is no need to define pKb separately from pKa,[32] but it is done here as often only pKb values can be found in the older literature.

For an hydrolyzed metal ion, Kb can also be defined as a stepwise dissociation constant

 
 

This is the reciprocal of an association constant for formation of the complex.

Basicity expressed as dissociation constant of conjugate acid edit

Because the relationship pKb = pKw − pKa holds only in aqueous solutions (though analogous relationships apply for other amphoteric solvents), subdisciplines of chemistry like organic chemistry that usually deal with nonaqueous solutions generally do not use pKb as a measure of basicity. Instead, the pKa of the conjugate acid, denoted by pKaH, is quoted when basicity needs to be quantified. For base B and its conjugate acid BH+ in equilibrium, this is defined as

 

A higher value for pKaH corresponds to a stronger base. For example, the values pKaH (C5H5N) = 5.25 and pKaH ((CH3CH2)3N) = 10.75 indicate that (CH3CH2)3N (triethylamine) is a stronger base than C5H5N (pyridine).

Amphoteric substances edit

An amphoteric substance is one that can act as an acid or as a base, depending on pH. Water (below) is amphoteric. Another example of an amphoteric molecule is the bicarbonate ion HCO3 that is the conjugate base of the carbonic acid molecule H2CO3 in the equilibrium

 

but also the conjugate acid of the carbonate ion CO2−3 in (the reverse of) the equilibrium

 

Carbonic acid equilibria are important for acid–base homeostasis in the human body.

An amino acid is also amphoteric with the added complication that the neutral molecule is subject to an internal acid–base equilibrium in which the basic amino group attracts and binds the proton from the acidic carboxyl group, forming a zwitterion.

 

At pH less than about 5 both the carboxylate group and the amino group are protonated. As pH increases the acid dissociates according to

 

At high pH a second dissociation may take place.

 

Thus the amino acid molecule is amphoteric because it may either be protonated or deprotonated.

Water self-ionization edit

The water molecule may either gain or lose a proton. It is said to be amphiprotic. The ionization equilibrium can be written

 

where in aqueous solution H+ denotes a solvated proton. Often this is written as the hydronium ion H3O+, but this formula is not exact because in fact there is solvation by more than one water molecule and species such as H5O+2, H7O+3, and H9O+4 are also present.[33]

The equilibrium constant is given by

 

With solutions in which the solute concentrations are not very high, the concentration [H2O] can be assumed to be constant, regardless of solute(s); this expression may then be replaced by

 

The self-ionization constant of water, Kw, is thus just a special case of an acid dissociation constant. A logarithmic form analogous to pKa may also be defined

 
pKw values for pure water at various temperatures[34]
T (°C) 0 5 10 15 20 25 30 35 40 45 50
pKw 14.943 14.734 14.535 14.346 14.167 13.997 13.830 13.680 13.535 13.396 13.262

These data can be modelled by a parabola with

 

From this equation, pKw = 14 at 24.87 °C. At that temperature both hydrogen and hydroxide ions have a concentration of 10−7 M.

Acidity in nonaqueous solutions edit

A solvent will be more likely to promote ionization of a dissolved acidic molecule in the following circumstances:[35]

  1. It is a protic solvent, capable of forming hydrogen bonds.
  2. It has a high donor number, making it a strong Lewis base.
  3. It has a high dielectric constant (relative permittivity), making it a good solvent for ionic species.

pKa values of organic compounds are often obtained using the aprotic solvents dimethyl sulfoxide (DMSO)[35] and acetonitrile (ACN).[36]

Solvent properties at 25 °C
Solvent Donor number[35] Dielectric constant[35]
Acetonitrile 14 37
Dimethylsulfoxide 30 47
Water 18 78

DMSO is widely used as an alternative to water because it has a lower dielectric constant than water, and is less polar and so dissolves non-polar, hydrophobic substances more easily. It has a measurable pKa range of about 1 to 30. Acetonitrile is less basic than DMSO, and, so, in general, acids are weaker and bases are stronger in this solvent. Some pKa values at 25 °C for acetonitrile (ACN)[37][38][39] and dimethyl sulfoxide (DMSO).[40] are shown in the following tables. Values for water are included for comparison.

pKa values of acids
HA ⇌ A + H+ ACN DMSO Water
p-Toluenesulfonic acid 8.5 0.9 Strong
2,4-Dinitrophenol 16.66 5.1 3.9
Benzoic acid 21.51 11.1 4.2
Acetic acid 23.51 12.6 4.756
Phenol 29.14 18.0 9.99
BH+ ⇌ B + H+ ACN DMSO Water
Pyrrolidine 19.56 10.8 11.4
Triethylamine 18.82 9.0 10.72
Proton sponge            18.62 7.5 12.1
Pyridine 12.53 3.4 5.2
Aniline 10.62 3.6 4.6

Ionization of acids is less in an acidic solvent than in water. For example, hydrogen chloride is a weak acid when dissolved in acetic acid. This is because acetic acid is a much weaker base than water.

 
 

Compare this reaction with what happens when acetic acid is dissolved in the more acidic solvent pure sulfuric acid:[41]

 
 
Dimerization of a carboxylic acid.

The unlikely geminal diol species CH3C(OH)+2 is stable in these environments. For aqueous solutions the pH scale is the most convenient acidity function.[42] Other acidity functions have been proposed for non-aqueous media, the most notable being the Hammett acidity function, H0, for superacid media and its modified version H for superbasic media.[43]

In aprotic solvents, oligomers, such as the well-known acetic acid dimer, may be formed by hydrogen bonding. An acid may also form hydrogen bonds to its conjugate base. This process, known as homoconjugation, has the effect of enhancing the acidity of acids, lowering their effective pKa values, by stabilizing the conjugate base. Homoconjugation enhances the proton-donating power of toluenesulfonic acid in acetonitrile solution by a factor of nearly 800.[44]

In aqueous solutions, homoconjugation does not occur, because water forms stronger hydrogen bonds to the conjugate base than does the acid.

Mixed solvents edit

 
pKa of acetic acid in dioxane/water mixtures. Data at 25 °C from Pine et al.[45]

When a compound has limited solubility in water it is common practice (in the pharmaceutical industry, for example) to determine pKa values in a solvent mixture such as water/dioxane or water/methanol, in which the compound is more soluble.[46] In the example shown at the right, the pKa value rises steeply with increasing percentage of dioxane as the dielectric constant of the mixture is decreasing.

A pKa value obtained in a mixed solvent cannot be used directly for aqueous solutions. The reason for this is that when the solvent is in its standard state its activity is defined as one. For example, the standard state of water:dioxane mixture with 9:1 mixing ratio is precisely that solvent mixture, with no added solutes. To obtain the pKa value for use with aqueous solutions it has to be extrapolated to zero co-solvent concentration from values obtained from various co-solvent mixtures.

These facts are obscured by the omission of the solvent from the expression that is normally used to define pKa, but pKa values obtained in a given mixed solvent can be compared to each other, giving relative acid strengths. The same is true of pKa values obtained in a particular non-aqueous solvent such a DMSO.

A universal, solvent-independent, scale for acid dissociation constants has not been developed, since there is no known way to compare the standard states of two different solvents.

Factors that affect pKa values edit

Pauling's second rule is that the value of the first pKa for acids of the formula XOm(OH)n depends primarily on the number of oxo groups m, and is approximately independent of the number of hydroxy groups n, and also of the central atom X. Approximate values of pKa are 8 for m = 0, 2 for m = 1, −3 for m = 2 and < −10 for m = 3.[28] Alternatively, various numerical formulas have been proposed including pKa = 8 − 5m (known as Bell's rule),[29][47] pKa = 7 − 5m,[30][48] or pKa = 9 − 7m.[29] The dependence on m correlates with the oxidation state of the central atom, X: the higher the oxidation state the stronger the oxyacid.

For example, pKa for HClO is 7.2, for HClO2 is 2.0, for HClO3 is −1 and HClO4 is a strong acid (pKa ≪ 0).[7] The increased acidity on adding an oxo group is due to stabilization of the conjugate base by delocalization of its negative charge over an additional oxygen atom.[47] This rule can help assign molecular structure: for example, phosphorous acid, having molecular formula H3PO3, has a pKa near 2, which suggested that the structure is HPO(OH)2, as later confirmed by NMR spectroscopy, and not P(OH)3, which would be expected to have a pKa near 8.[48]

 
pKa values for acetic, chloroacetic, dichloroacetic and trichloroacetic acids.

Inductive effects and mesomeric effects affect the pKa values. A simple example is provided by the effect of replacing the hydrogen atoms in acetic acid by the more electronegative chlorine atom. The electron-withdrawing effect of the substituent makes ionisation easier, so successive pKa values decrease in the series 4.7, 2.8, 1.4, and 0.7 when 0, 1, 2, or 3 chlorine atoms are present.[49] The Hammett equation, provides a general expression for the effect of substituents.[50]

log(Ka) = log(K0
a
) + ρσ.

Ka is the dissociation constant of a substituted compound, K0
a
is the dissociation constant when the substituent is hydrogen, ρ is a property of the unsubstituted compound and σ has a particular value for each substituent. A plot of log(Ka) against σ is a straight line with intercept log(K0
a
) and slope ρ. This is an example of a linear free energy relationship as log(Ka) is proportional to the standard free energy change. Hammett originally[51] formulated the relationship with data from benzoic acid with different substituents in the ortho- and para- positions: some numerical values are in Hammett equation. This and other studies allowed substituents to be ordered according to their electron-withdrawing or electron-releasing power, and to distinguish between inductive and mesomeric effects.[52][53]

Alcohols do not normally behave as acids in water, but the presence of a double bond adjacent to the OH group can substantially decrease the pKa by the mechanism of keto–enol tautomerism. Ascorbic acid is an example of this effect. The diketone 2,4-pentanedione (acetylacetone) is also a weak acid because of the keto–enol equilibrium. In aromatic compounds, such as phenol, which have an OH substituent, conjugation with the aromatic ring as a whole greatly increases the stability of the deprotonated form.

Structural effects can also be important. The difference between fumaric acid and maleic acid is a classic example. Fumaric acid is (E)-1,4-but-2-enedioic acid, a trans isomer, whereas maleic acid is the corresponding cis isomer, i.e. (Z)-1,4-but-2-enedioic acid (see cis-trans isomerism). Fumaric acid has pKa values of approximately 3.0 and 4.5. By contrast, maleic acid has pKa values of approximately 1.5 and 6.5. The reason for this large difference is that when one proton is removed from the cis isomer (maleic acid) a strong intramolecular hydrogen bond is formed with the nearby remaining carboxyl group. This favors the formation of the maleate H+, and it opposes the removal of the second proton from that species. In the trans isomer, the two carboxyl groups are always far apart, so hydrogen bonding is not observed.[54]

 
Proton sponge

Proton sponge, 1,8-bis(dimethylamino)naphthalene, has a pKa value of 12.1. It is one of the strongest amine bases known. The high basicity is attributed to the relief of strain upon protonation and strong internal hydrogen bonding.[55][56]

Effects of the solvent and solvation should be mentioned also in this section. It turns out, these influences are more subtle than that of a dielectric medium mentioned above. For example, the expected (by electronic effects of methyl substituents) and observed in gas phase order of basicity of methylamines, Me3N > Me2NH > MeNH2 > NH3, is changed by water to Me2NH > MeNH2 > Me3N > NH3. Neutral methylamine molecules are hydrogen-bonded to water molecules mainly through one acceptor, N–HOH, interaction and only occasionally just one more donor bond, NH–OH2. Hence, methylamines are stabilized to about the same extent by hydration, regardless of the number of methyl groups. In stark contrast, corresponding methylammonium cations always utilize all the available protons for donor NH–OH2 bonding. Relative stabilization of methylammonium ions thus decreases with the number of methyl groups explaining the order of water basicity of methylamines.[4]

Thermodynamics edit

An equilibrium constant is related to the standard Gibbs energy change for the reaction, so for an acid dissociation constant

 .

R is the gas constant and T is the absolute temperature. Note that pKa = −log(Ka) and 2.303 ≈ ln(10). At 25 °C, ΔG in kJ·mol−1 ≈ 5.708 pKa (1 kJ·mol−1 = 1000 joules per mole). Free energy is made up of an enthalpy term and an entropy term.[11]

 

The standard enthalpy change can be determined by calorimetry or by using the van 't Hoff equation, though the calorimetric method is preferable. When both the standard enthalpy change and acid dissociation constant have been determined, the standard entropy change is easily calculated from the equation above. In the following table, the entropy terms are calculated from the experimental values of pKa and ΔH. The data were critically selected and refer to 25 °C and zero ionic strength, in water.[11]

Acids
Compound Equilibrium pKa ΔG (kJ·mol−1)[d] ΔH (kJ·mol−1) TΔS (kJ·mol−1)[e]
HA = Acetic acid HA ⇌ H+ + A 4.756 27.147 −0.41 27.56
H2A+ = GlycineH+ H2A+ ⇌ HA + H+ 2.351 13.420 4.00 9.419
HA ⇌ H+ + A 9.78 55.825 44.20 11.6
H2A = Maleic acid H2A ⇌ HA + H+ 1.92 10.76 1.10 9.85
HA ⇌ H+ + A2− 6.27 35.79 −3.60 39.4
H3A = Citric acid H3A ⇌ H2A + H+ 3.128 17.855 4.07 13.78
H2A ⇌ HA2− + H+ 4.76 27.176 2.23 24.9
HA2− ⇌ A3− + H+ 6.40 36.509 −3.38 39.9
H3A = Boric acid H3A ⇌ H2A + H+ 9.237 52.725 13.80 38.92
H3A = Phosphoric acid H3A ⇌ H2A + H+ 2.148 12.261 −8.00 20.26
H2A ⇌ HA2− + H+ 7.20 41.087 3.60 37.5
HA2− ⇌ A3− + H+ 12.35 80.49 16.00 54.49
HA = Hydrogen sulfate HA ⇌ A2− + H+ 1.99 11.36 −22.40 33.74
H2A = Oxalic acid H2A ⇌ HA + H+ 1.27 7.27 −3.90 11.15
HA ⇌ A2− + H+ 4.266 24.351 −7.00 31.35
  1. ^ The hydrogen ion does not exist as such in solution. It combines with a solvent molecule; when the solvent is water a hydronium ion is formed: H+ + H2O → H3O+. This reaction is quantitative and hence can be ignored in the context of chemical equilibrium.
  2. ^ It is common practice to quote pK values rather than K values. pK = −log10 K. pKa is often referred to as an acid dissociation constant, but this is, strictly speaking, incorrect as pKa is the cologarithm of the dissociation constant.
  3. ^ It is implicit in this definition that the quotient of activity coefficients,   is a constant with a value of 1 under a given set of experimental conditions.
  4. ^ ΔG ≈ 2.303RTpKa
  5. ^ Computed here, from ΔH and ΔG values supplied in the citation, using TΔS = ΔG − ΔH
Conjugate acids of bases
Compound Equilibrium pKa ΔH (kJ·mol−1) TΔS (kJ·mol−1)
B = Ammonia HB+ ⇌ B + H+ 9.245 51.95 0.8205
B = Methylamine HB+ ⇌ B + H+ 10.645 55.34 5.422
B = Triethylamine HB+ ⇌ B + H+ 10.72 43.13 18.06

The first point to note is that, when pKa is positive, the standard free energy change for the dissociation reaction is also positive. Second, some reactions are exothermic and some are endothermic, but, when ΔH is negative TΔS is the dominant factor, which determines that ΔG is positive. Last, the entropy contribution is always unfavourable (ΔS < 0) in these reactions. Ions in aqueous solution tend to orient the surrounding water molecules, which orders the solution and decreases the entropy. The contribution of an ion to the entropy is the partial molar entropy which is often negative, especially for small or highly charged ions.[57] The ionization of a neutral acid involves formation of two ions so that the entropy decreases (ΔS < 0). On the second ionization of the same acid, there are now three ions and the anion has a charge, so the entropy again decreases.

Note that the standard free energy change for the reaction is for the changes from the reactants in their standard states to the products in their standard states. The free energy change at equilibrium is zero since the chemical potentials of reactants and products are equal at equilibrium.

Experimental determination edit

 
A calculated titration curve of oxalic acid titrated with a solution of sodium hydroxide

The experimental determination of pKa values is commonly performed by means of titrations, in a medium of high ionic strength and at constant temperature.[58] A typical procedure would be as follows. A solution of the compound in the medium is acidified with a strong acid to the point where the compound is fully protonated. The solution is then titrated with a strong base until all the protons have been removed. At each point in the titration pH is measured using a glass electrode and a pH meter. The equilibrium constants are found by fitting calculated pH values to the observed values, using the method of least squares.[59]

The total volume of added strong base should be small compared to the initial volume of titrand solution in order to keep the ionic strength nearly constant. This will ensure that pKa remains invariant during the titration.

A calculated titration curve for oxalic acid is shown at the right. Oxalic acid has pKa values of 1.27 and 4.27. Therefore, the buffer regions will be centered at about pH 1.3 and pH 4.3. The buffer regions carry the information necessary to get the pKa values as the concentrations of acid and conjugate base change along a buffer region.

Between the two buffer regions there is an end-point, or equivalence point, at about pH 3. This end-point is not sharp and is typical of a diprotic acid whose buffer regions overlap by a small amount: pKa2 − pKa1 is about three in this example. (If the difference in pK values were about two or less, the end-point would not be noticeable.) The second end-point begins at about pH 6.3 and is sharp. This indicates that all the protons have been removed. When this is so, the solution is not buffered and the pH rises steeply on addition of a small amount of strong base. However, the pH does not continue to rise indefinitely. A new buffer region begins at about pH 11 (pKw − 3), which is where self-ionization of water becomes important.

It is very difficult to measure pH values of less than two in aqueous solution with a glass electrode, because the Nernst equation breaks down at such low pH values. To determine pK values of less than about 2 or more than about 11 spectrophotometric[60][61] or NMR[62][63] measurements may be used instead of, or combined with, pH measurements.

When the glass electrode cannot be employed, as with non-aqueous solutions, spectrophotometric methods are frequently used.[38] These may involve absorbance or fluorescence measurements. In both cases the measured quantity is assumed to be proportional to the sum of contributions from each photo-active species; with absorbance measurements the Beer–Lambert law is assumed to apply.

Isothermal titration calorimetry (ITC) may be used to determine both a pK value and the corresponding standard enthalpy for acid dissociation.[64] Software to perform the calculations is supplied by the instrument manufacturers for simple systems.

Aqueous solutions with normal water cannot be used for 1H NMR measurements but heavy water, D2O, must be used instead. 13C NMR data, however, can be used with normal water and 1H NMR spectra can be used with non-aqueous media. The quantities measured with NMR are time-averaged chemical shifts, as proton exchange is fast on the NMR time-scale. Other chemical shifts, such as those of 31P can be measured.

Micro-constants edit

 
Cysteine

For some polyprotic acids, dissociation (or association) occurs at more than one nonequivalent site,[4] and the observed macroscopic equilibrium constant or macroconstant is a combination of microconstants involving distinct species. When one reactant forms two products in parallel, the macroconstant is a sum of two microconstants,   This is true for example for the deprotonation of the amino acid cysteine, which exists in solution as a neutral zwitterion HS−CH2−CH(NH+3)−COO. The two microconstants represent deprotonation either at sulphur or at nitrogen, and the macroconstant sum here is the acid dissociation constant  [65]

 
Spermine

Similarly, a base such as spermine has more than one site where protonation can occur. For example, monoprotonation can occur at a terminal −NH2 group or at internal −NH− groups. The Kb values for dissociation of spermine protonated at one or other of the sites are examples of micro-constants. They cannot be determined directly by means of pH, absorbance, fluorescence or NMR measurements; a measured Kb value is the sum of the K values for the micro-reactions.

 

Nevertheless, the site of protonation is very important for biological function, so mathematical methods have been developed for the determination of micro-constants.[66]

When two reactants form a single product in parallel, the macroconstant  [65] For example, the abovementioned equilibrium for spermine may be considered in terms of Ka values of two tautomeric conjugate acids, with macroconstant In this case   This is equivalent to the preceding expression since   is proportional to  

When a reactant undergoes two reactions in series, the macroconstant for the combined reaction is the product of the microconstant for the two steps. For example, the abovementioned cysteine zwitterion can lose two protons, one from sulphur and one from nitrogen, and the overall macroconstant for losing two protons is the product of two dissociation constants  [65] This can also be written in terms of logarithmic constants as  

Applications and significance edit

A knowledge of pKa values is important for the quantitative treatment of systems involving acid–base equilibria in solution. Many applications exist in biochemistry; for example, the pKa values of proteins and amino acid side chains are of major importance for the activity of enzymes and the stability of proteins.[67] Protein pKa values cannot always be measured directly, but may be calculated using theoretical methods. Buffer solutions are used extensively to provide solutions at or near the physiological pH for the study of biochemical reactions;[68] the design of these solutions depends on a knowledge of the pKa values of their components. Important buffer solutions include MOPS, which provides a solution with pH 7.2, and tricine, which is used in gel electrophoresis.[69][70] Buffering is an essential part of acid base physiology including acid–base homeostasis,[71] and is key to understanding disorders such as acid–base disorder.[72][73][74] The isoelectric point of a given molecule is a function of its pK values, so different molecules have different isoelectric points. This permits a technique called isoelectric focusing,[75] which is used for separation of proteins by 2-D gel polyacrylamide gel electrophoresis.

Buffer solutions also play a key role in analytical chemistry. They are used whenever there is a need to fix the pH of a solution at a particular value. Compared with an aqueous solution, the pH of a buffer solution is relatively insensitive to the addition of a small amount of strong acid or strong base. The buffer capacity[76] of a simple buffer solution is largest when pH = pKa. In acid–base extraction, the efficiency of extraction of a compound into an organic phase, such as an ether, can be optimised by adjusting the pH of the aqueous phase using an appropriate buffer. At the optimum pH, the concentration of the electrically neutral species is maximised; such a species is more soluble in organic solvents having a low dielectric constant than it is in water. This technique is used for the purification of weak acids and bases.[77]

A pH indicator is a weak acid or weak base that changes colour in the transition pH range, which is approximately pKa ± 1. The design of a universal indicator requires a mixture of indicators whose adjacent pKa values differ by about two, so that their transition pH ranges just overlap.

In pharmacology, ionization of a compound alters its physical behaviour and macro properties such as solubility and lipophilicity, log p). For example, ionization of any compound will increase the solubility in water, but decrease the lipophilicity. This is exploited in drug development to increase the concentration of a compound in the blood by adjusting the pKa of an ionizable group.[78]

Knowledge of pKa values is important for the understanding of coordination complexes, which are formed by the interaction of a metal ion, Mm+, acting as a Lewis acid, with a ligand, L, acting as a Lewis base. However, the ligand may also undergo protonation reactions, so the formation of a complex in aqueous solution could be represented symbolically by the reaction

 

To determine the equilibrium constant for this reaction, in which the ligand loses a proton, the pKa of the protonated ligand must be known. In practice, the ligand may be polyprotic; for example EDTA4− can accept four protons; in that case, all pKa values must be known. In addition, the metal ion is subject to hydrolysis, that is, it behaves as a weak acid, so the pK values for the hydrolysis reactions must also be known.[79]

Assessing the hazard associated with an acid or base may require a knowledge of pKa values.[80] For example, hydrogen cyanide is a very toxic gas, because the cyanide ion inhibits the iron-containing enzyme cytochrome c oxidase. Hydrogen cyanide is a weak acid in aqueous solution with a pKa of about 9. In strongly alkaline solutions, above pH 11, say, it follows that sodium cyanide is "fully dissociated" so the hazard due to the hydrogen cyanide gas is much reduced. An acidic solution, on the other hand, is very hazardous because all the cyanide is in its acid form. Ingestion of cyanide by mouth is potentially fatal, independently of pH, because of the reaction with cytochrome c oxidase.

In environmental science acid–base equilibria are important for lakes[81] and rivers;[82][83] for example, humic acids are important components of natural waters. Another example occurs in chemical oceanography:[84] in order to quantify the solubility of iron(III) in seawater at various salinities, the pKa values for the formation of the iron(III) hydrolysis products Fe(OH)2+, Fe(OH)+2 and Fe(OH)3 were determined, along with the solubility product of iron hydroxide.[85]

Values for common substances edit

There are multiple techniques to determine the pKa of a chemical, leading to some discrepancies between different sources. Well measured values are typically within 0.1 units of each other. Data presented here were taken at 25 °C in water.[7][86] More values can be found in the Thermodynamics section, above. A table of pKa of carbon acids, measured in DMSO, can be found on the page on carbanions.

Chemical Equilibrium pKa
BH = Adenine BH ⇌ B + H+ 4.17
BH+
2
⇌ BH + H+
9.65
H3A = Arsenic acid H3A ⇌ H2A + H+ 2.22
H2A ⇌ HA2− + H+ 6.98
HA2− ⇌ A3− + H+ 11.53
HA = Benzoic acid HA ⇌ H+ + A 4.204
HA = Butyric acid HA ⇌ H+ + A 4.82
H2A = Chromic acid H2A ⇌ HA + H+ 0.98
HA ⇌ A2− + H+ 6.5
B = Codeine BH+ ⇌ B + H+ 8.17
HA = Cresol HA ⇌ H+ + A 10.29
HA = Formic acid HA ⇌ H+ + A 3.751
HA = Hydrofluoric acid HA ⇌ H+ + A 3.17
HA = Hydrocyanic acid HA ⇌ H+ + A 9.21
HA = Hydrogen selenide HA ⇌ H+ + A 3.89
HA = Hydrogen peroxide (90%) HA ⇌ H+ + A 11.7
HA = Lactic acid HA ⇌ H+ + A 3.86
HA = Propionic acid HA ⇌ H+ + A 4.87
HA = Phenol HA ⇌ H+ + A 9.99
H2A = L-(+)-Ascorbic Acid H2A ⇌ HA + H+ 4.17
HA ⇌ A2− + H+ 11.57

See also edit

Notes edit

References edit

  1. ^ Whitten, Kenneth W.; Gailey, Kenneth D.; Davis, Raymond E. (1992). General Chemistry (4th ed.). Saunders College Publishing. p. 660. ISBN 0-03-072373-6.
  2. ^ Petrucci, Ralph H.; Harwood, William S.; Herring, F. Geoffrey (2002). General Chemistry (8th ed.). Prentice Hall. pp. 667–8. ISBN 0-13-014329-4.
  3. ^ Perrin DD, Dempsey B, Serjeant EP (1981). "Chapter 3: Methods of pKa Prediction". pKa Prediction for Organic Acids and Bases. (secondary). London: Chapman & Hall. pp. 21–26. doi:10.1007/978-94-009-5883-8. ISBN 978-0-412-22190-3.
  4. ^ a b c Fraczkiewicz R (2013). "In Silico Prediction of Ionization". In Reedijk J (ed.). Reference Module in Chemistry, Molecular Sciences and Chemical Engineering. (secondary). Reference Module in Chemistry, Molecular Sciences and Chemical Engineering [Online]. Vol. 5. Amsterdam, the Netherlands: Elsevier. doi:10.1016/B978-0-12-409547-2.02610-X. ISBN 9780124095472.
  5. ^ Miessler, Gary L.; Tarr, Donald A. (1991). Inorganic Chemistry (2nd ed.). Prentice Hall. ISBN 0-13-465659-8. Chapter 6: Acid–Base and Donor–Acceptor Chemistry
  6. ^ a b Bell, R.P. (1973). The Proton in Chemistry (2nd ed.). London: Chapman & Hall. ISBN 0-8014-0803-2. Includes discussion of many organic Brønsted acids.
  7. ^ a b c Shriver, D.F; Atkins, P.W. (1999). Inorganic Chemistry (3rd ed.). Oxford: Oxford University Press. ISBN 0-19-850331-8. Chapter 5: Acids and Bases
  8. ^ Housecroft, C. E.; Sharpe, A. G. (2008). Inorganic Chemistry (3rd ed.). Prentice Hall. ISBN 978-0-13-175553-6. Chapter 6: Acids, Bases and Ions in Aqueous Solution
  9. ^ Headrick, J.M.; Diken, E.G.; Walters, R. S.; Hammer, N. I.; Christie, R.A.; Cui, J.; Myshakin, E.M.; Duncan, M.A.; Johnson, M.A.; Jordan, K.D. (2005). "Spectral Signatures of Hydrated Proton Vibrations in Water Clusters". Science. 308 (5729): 1765–69. Bibcode:2005Sci...308.1765H. doi:10.1126/science.1113094. PMID 15961665. S2CID 40852810.
  10. ^ Smiechowski, M.; Stangret, J. (2006). "Proton hydration in aqueous solution: Fourier transform infrared studies of HDO spectra". J. Chem. Phys. 125 (20): 204508–204522. Bibcode:2006JChPh.125t4508S. doi:10.1063/1.2374891. PMID 17144716.
  11. ^ a b c Goldberg, R.; Kishore, N.; Lennen, R. (2002). (PDF). J. Phys. Chem. Ref. Data. 31 (2): 231–370. Bibcode:2002JPCRD..31..231G. doi:10.1063/1.1416902. Archived from the original (PDF) on 2008-10-06.
  12. ^ Jolly, William L. (1984). Modern Inorganic Chemistry. McGraw-Hill. pp. 198. ISBN 978-0-07-032760-3.
  13. ^ Burgess, J. (1978). Metal Ions in Solution. Ellis Horwood. ISBN 0-85312-027-7. Section 9.1 "Acidity of Solvated Cations" lists many pKa values.
  14. ^ Petrucci, R.H.; Harwood, R.S.; Herring, F.G. (2002). General Chemistry (8th ed.). Prentice Hall. ISBN 0-13-014329-4. p.698
  15. ^ a b Rossotti, F.J.C.; Rossotti, H. (1961). The Determination of Stability Constants. McGraw–Hill. Chapter 2: Activity and Concentration Quotients pp 5-10
  16. ^ "Project: Ionic Strength Corrections for Stability Constants". International Union of Pure and Applied Chemistry. Retrieved 2019-03-28.
  17. ^ Rossotti, Francis J. C; Rozotti, Hazel (1961). The determination of stability constants : and other equilibrium constants in solution. New York: McGraw-Hill. pp. 5–10. ISBN 9781013909146. Archived from the original on 7 February 2020.
  18. ^ Atkins, P.W.; de Paula, J. (2006). Physical Chemistry. Oxford University Press. ISBN 0-19-870072-5. Section 7.4: The Response of Equilibria to Temperature
  19. ^ Petrucci, Ralph H.; Harwood, William S.; Herring, F. Geoffrey (2002). General chemistry: principles and modern applications (8th ed.). Prentice Hall. p. 633. ISBN 0-13-014329-4. Are you wondering... How using activities makes the equilibrium constant dimensionless?
  20. ^ Shriver, D.F; Atkins, P.W. (1999). Inorganic Chemistry (3rd ed.). Oxford University Press. ISBN 0-19-850331-8. Sec. 5.1c Strong and weak acids and bases
  21. ^ Porterfield, William W. (1984). Inorganic Chemistry. Addison-Wesley. p. 260. ISBN 0-201-05660-7.
  22. ^ a b Shriver, D.F; Atkins, P.W. (1999). Inorganic Chemistry (3rd ed.). Oxford University Press. ISBN 0-19-850331-8. Sec. 5.2 Solvent leveling
  23. ^ Levanov, A. V.; Isaikina, O. Ya.; Lunin, V. V. (2017). "Dissociation constant of nitric acid". Russian Journal of Physical Chemistry A. 91 (7): 1221–1228. Bibcode:2017RJPCA..91.1221L. doi:10.1134/S0036024417070196. S2CID 104093297.
  24. ^ Trummal, Aleksander; Lipping, Lauri; Kaljurand, Ivari; Koppel, Ilmar A.; Leito, Ivo (2016). "Acidity of Strong Acids in Water and Dimethyl Sulfoxide". The Journal of Physical Chemistry A. 120 (20): 3663–3669. Bibcode:2016JPCA..120.3663T. doi:10.1021/acs.jpca.6b02253. PMID 27115918. S2CID 29697201.
  25. ^ Mehta, Akul (22 October 2012). "Henderson–Hasselbalch Equation: Derivation of pKa and pKb". PharmaXChange. Retrieved 16 November 2014.
  26. ^ The values are for 25 °C and 0 ionic strength – Powell, Kipton J.; Brown, Paul L.; Byrne, Robert H.; Gajda, Tamás; Hefter, Glenn; Sjöberg, Staffan; Wanner, Hans (2005). "Chemical speciation of environmentally significant heavy metals with inorganic ligands. Part 1: The Hg2+ – Cl, OH, CO32-, SO42-, and PO43- aqueous systems". Pure Appl. Chem. 77 (4): 739–800. doi:10.1351/pac200577040739.
  27. ^ Brown, T.E.; Lemay, H.E.; Bursten, B.E.; Murphy, C.; Woodward, P. (2008). Chemistry: The Central Science (11th ed.). New York: Prentice-Hall. p. 689. ISBN 978-0-13-600617-6.
  28. ^ a b Greenwood, N.N.; Earnshaw, A. (1997). Chemistry of the Elements (2nd ed.). Oxford: Butterworth-Heinemann. p. 50. ISBN 0-7506-3365-4.
  29. ^ a b c Miessler, Gary L.; Tarr Donald A. (1999). Inorganic Chemistry (2nd ed.). Prentice Hall. p. 164. ISBN 0-13-465659-8.
  30. ^ a b Huheey, James E. (1983). Inorganic Chemistry (3rd ed.). Harper & Row. p. 297. ISBN 0-06-042987-9.
  31. ^ Lide, D.R. (2004). CRC Handbook of Chemistry and Physics, Student Edition (84th ed.). CRC Press. ISBN 0-8493-0597-7. Section D–152
  32. ^ Skoog, Douglas A.; West, Donald M.; Holler, F. James; Crouch, Stanley R. (2014). Fundamentals of Analytical Chemistry (9th ed.). Brooks/Cole. p. 212. ISBN 978-0-495-55828-6.
  33. ^ Housecroft, C. E.; Sharpe, A. G. (2004). Inorganic Chemistry (2nd ed.). Prentice Hall. p. 163. ISBN 978-0-13-039913-7.
  34. ^ Harned, H.S.; Owen, B.B (1958). The Physical Chemistry of Electrolytic Solutions. New York: Reinhold Publishing Corp. pp. 634–649, 752–754.
  35. ^ a b c d Loudon, G. Marc (2005), Organic Chemistry (4th ed.), New York: Oxford University Press, pp. 317–318, ISBN 0-19-511999-1
  36. ^ March, J.; Smith, M. (2007). Advanced Organic Chemistry (6th ed.). New York: John Wiley & Sons. ISBN 978-0-471-72091-1. Chapter 8: Acids and Bases
  37. ^ Kütt, A.; Movchun, V.; Rodima, T; Dansauer, T.; Rusanov, E.B.; Leito, I.; Kaljurand, I.; Koppel, J.; Pihl, V.; Koppel, I.; Ovsjannikov, G.; Toom, L.; Mishima, M.; Medebielle, M.; Lork, E.; Röschenthaler, G-V.; Koppel, I.A.; Kolomeitsev, A.A. (2008). "Pentakis(trifluoromethyl)phenyl, a Sterically Crowded and Electron-withdrawing Group: Synthesis and Acidity of Pentakis(trifluoromethyl)benzene, -toluene, -phenol, and -aniline". J. Org. Chem. 73 (7): 2607–2620. doi:10.1021/jo702513w. PMID 18324831.
  38. ^ a b Kütt, A.; Leito, I.; Kaljurand, I.; Sooväli, L.; Vlasov, V.M.; Yagupolskii, L.M.; Koppel, I.A. (2006). "A Comprehensive Self-Consistent Spectrophotometric Acidity Scale of Neutral Brønsted Acids in Acetonitrile". J. Org. Chem. 71 (7): 2829–2838. doi:10.1021/jo060031y. PMID 16555839. S2CID 8596886.
  39. ^ Kaljurand, I.; Kütt, A.; Sooväli, L.; Rodima, T.; Mäemets, V.; Leito, I; Koppel, I.A. (2005). "Extension of the Self-Consistent Spectrophotometric Basicity Scale in Acetonitrile to a Full Span of 28 pKa Units: Unification of Different Basicity Scales". J. Org. Chem. 70 (3): 1019–1028. doi:10.1021/jo048252w. PMID 15675863.
  40. ^ . Archived from the original on 9 October 2008. Retrieved 2008-11-02.
  41. ^ Housecroft, C. E.; Sharpe, A. G. (2008). Inorganic Chemistry (3rd ed.). Prentice Hall. ISBN 978-0-13-175553-6. Chapter 8: Non-Aqueous Media
  42. ^ Rochester, C.H. (1970). Acidity Functions. Academic Press. ISBN 0-12-590850-4.
  43. ^ Olah, G.A; Prakash, S; Sommer, J (1985). Superacids. New York: Wiley Interscience. ISBN 0-471-88469-3.
  44. ^ Coetzee, J.F.; Padmanabhan, G.R. (1965). "Proton Acceptor Power and Homoconjugation of Mono- and Diamines". J. Am. Chem. Soc. 87 (22): 5005–5010. doi:10.1021/ja00950a006.
  45. ^ Pine, S.H.; Hendrickson, J.B.; Cram, D.J.; Hammond, G.S. (1980). Organic chemistry. McGraw–Hill. p. 203. ISBN 0-07-050115-7.
  46. ^ Box, K.J.; Völgyi, G.; Ruiz, R.; Comer, J.E.; Takács-Novák, K.; Bosch, E.; Ràfols, C.; Rosés, M. (2007). "Physicochemical Properties of a New Multicomponent Cosolvent System for the pKa Determination of Poorly Soluble Pharmaceutical Compounds". Helv. Chim. Acta. 90 (8): 1538–1553. doi:10.1002/hlca.200790161.
  47. ^ a b Housecroft, Catherine E.; Sharpe, Alan G. (2005). Inorganic chemistry (2nd ed.). Harlow, U.K.: Pearson Prentice Hall. pp. 170–171. ISBN 0-13-039913-2.
  48. ^ a b Douglas B., McDaniel D.H. and Alexander J.J. Concepts and Models of Inorganic Chemistry (2nd ed. Wiley 1983) p.526 ISBN 0-471-21984-3
  49. ^ Pauling, L. (1960). The nature of the chemical bond and the structure of molecules and crystals; an introduction to modern structural chemistry (3rd ed.). Ithaca (NY): Cornell University Press. p. 277. ISBN 0-8014-0333-2.
  50. ^ Pine, S.H.; Hendrickson, J.B.; Cram, D.J.; Hammond, G.S. (1980). Organic Chemistry. McGraw–Hill. ISBN 0-07-050115-7. Section 13-3: Quantitative Correlations of Substituent Effects (Part B) – The Hammett Equation
  51. ^ Hammett, L.P. (1937). "The Effect of Structure upon the Reactions of Organic Compounds. Benzene Derivatives". J. Am. Chem. Soc. 59 (1): 96–103. doi:10.1021/ja01280a022.
  52. ^ Hansch, C.; Leo, A.; Taft, R. W. (1991). "A Survey of Hammett Substituent Constants and Resonance and Field Parameters". Chem. Rev. 91 (2): 165–195. doi:10.1021/cr00002a004. S2CID 97583278.
  53. ^ Shorter, J (1997). "Compilation and critical evaluation of structure-reactivity parameters and equations: Part 2. Extension of the Hammett σ scale through data for the ionization of substituted benzoic acids in aqueous solvents at 25 °C (Technical Report)". Pure and Applied Chemistry. 69 (12): 2497–2510. doi:10.1351/pac199769122497. S2CID 98814841.
  54. ^ Pine, S.H.; Hendrickson, J.B.; Cram, D.J.; Hammond, G.S. (1980). Organic chemistry. McGraw–Hill. ISBN 0-07-050115-7. Section 6-2: Structural Effects on Acidity and Basicity
  55. ^ Alder, R.W.; Bowman, P.S.; Steele, W.R.S.; Winterman, D.R. (1968). "The Remarkable Basicity of 1,8-bis(dimethylamino)naphthalene". Chem. Commun. (13): 723–724. doi:10.1039/C19680000723.
  56. ^ Alder, R.W. (1989). "Strain Effects on Amine Basicities". Chem. Rev. 89 (5): 1215–1223. doi:10.1021/cr00095a015.
  57. ^ Atkins, Peter William; De Paula, Julio (2006). Atkins' physical chemistry. New York: W H Freeman. p. 94. ISBN 978-0-7167-7433-4.
  58. ^ Martell, A.E.; Motekaitis, R.J. (1992). Determination and Use of Stability Constants. Wiley. ISBN 0-471-18817-4. Chapter 4: Experimental Procedure for Potentiometric pH Measurement of Metal Complex Equilibria
  59. ^ Leggett, D.J. (1985). Computational Methods for the Determination of Formation Constants. Plenum. ISBN 0-306-41957-2.
  60. ^ Allen, R.I.; Box, K.J.; Comer, J.E.A.; Peake, C.; Tam, K.Y. (1998). "Multiwavelength Spectrophotometric Determination of Acid Dissociation Constants of Ionizable Drugs". J. Pharm. Biomed. Anal. 17 (4–5): 699–712. doi:10.1016/S0731-7085(98)00010-7. PMID 9682153.
  61. ^ Box, K.J.; Donkor, R.E.; Jupp, P.A.; Leader, I.P.; Trew, D.F.; Turner, C.H. (2008). "The Chemistry of Multi-Protic Drugs Part 1: A Potentiometric, Multi-Wavelength UV and NMR pH Titrimetric Study of the Micro-Speciation of SKI-606". J. Pharm. Biomed. Anal. 47 (2): 303–311. doi:10.1016/j.jpba.2008.01.015. PMID 18314291.
  62. ^ Popov, K.; Ronkkomaki, H.; Lajunen, L.H.J. (2006). "Guidelines for NMR easurements for Determination of High and Low pKa Values" (PDF). Pure Appl. Chem. 78 (3): 663–675. doi:10.1351/pac200678030663. S2CID 4823180.
  63. ^ Szakács, Z.; Hägele, G. (2004). "Accurate Determination of Low pK Values by 1H NMR Titration". Talanta. 62 (4): 819–825. doi:10.1016/j.talanta.2003.10.007. PMID 18969368.
  64. ^ Feig, Andrew L., ed. (2016). "Methods in Enzymology". Calorimetry. Elsevier. 567: 2–493. ISSN 0076-6879.
  65. ^ a b c Splittgerber, A. G.; Chinander, L.L. (1 February 1988). "The spectrum of a dissociation intermediate of cysteine: a biophysical chemistry experiment". Journal of Chemical Education. 65 (2): 167. Bibcode:1988JChEd..65..167S. doi:10.1021/ed065p167.
  66. ^ Frassineti, C.; Alderighi, L; Gans, P; Sabatini, A; Vacca, A; Ghelli, S. (2003). "Determination of Protonation Constants of Some Fluorinated Polyamines by Means of 13C NMR Data Processed by the New Computer Program HypNMR2000. Protonation Sequence in Polyamines". Anal. Bioanal. Chem. 376 (7): 1041–1052. doi:10.1007/s00216-003-2020-0. PMID 12845401. S2CID 14533024.
  67. ^ Onufriev, A.; Case, D.A; Ullmann G.M. (2001). "A Novel View of pH Titration in Biomolecules". Biochemistry. 40 (12): 3413–3419. doi:10.1021/bi002740q. PMID 11297406.
  68. ^ Good, N.E.; Winget, G.D.; Winter, W.; Connolly, T.N.; Izawa, S.; Singh, R.M.M. (1966). "Hydrogen Ion Buffers for Biological Research". Biochemistry. 5 (2): 467–477. doi:10.1021/bi00866a011. PMID 5942950.
  69. ^ Dunn, M.J. (1993). Gel Electrophoresis: Proteins. Bios Scientific Publishers. ISBN 1-872748-21-X.
  70. ^ Martin, R. (1996). Gel Electrophoresis: Nucleic Acids. Bios Scientific Publishers. ISBN 1-872748-28-7.
  71. ^ Brenner, B.M.; Stein, J.H., eds. (1979). Acid–Base and Potassium Homeostasis. Churchill Livingstone. ISBN 0-443-08017-8.
  72. ^ Scorpio, R. (2000). Fundamentals of Acids, Bases, Buffers & Their Application to Biochemical Systems. Kendall/Hunt Pub. Co. ISBN 0-7872-7374-0.
  73. ^ Beynon, R.J.; Easterby, J.S. (1996). Buffer Solutions: The Basics. Oxford: Oxford University Press. ISBN 0-19-963442-4.
  74. ^ Perrin, D.D.; Dempsey, B. (1974). Buffers for pH and Metal Ion Control. London: Chapman & Hall. ISBN 0-412-11700-2.
  75. ^ Garfin, D.; Ahuja, S., eds. (2005). Handbook of Isoelectric Focusing and Proteomics. Vol. 7. Elsevier. ISBN 0-12-088752-5.
  76. ^ Hulanicki, A. (1987). Reactions of Acids and Bases in Analytical Chemistry. Masson, M.R. (translation editor). Horwood. ISBN 0-85312-330-6.
  77. ^ Eyal, A.M (1997). "Acid Extraction by Acid–Base-Coupled Extractants". Ion Exchange and Solvent Extraction: A Series of Advances. 13: 31–94.
  78. ^ Avdeef, A. (2003). Absorption and Drug Development: Solubility, Permeability, and Charge State. New York: Wiley. ISBN 0-471-42365-3.
  79. ^ Beck, M.T.; Nagypál, I. (1990). Chemistry of Complex Equilibria. Horwood. ISBN 0-85312-143-5.
  80. ^ van Leeuwen, C.J.; Hermens, L. M. (1995). Risk Assessment of Chemicals: An Introduction. Springer. pp. 254–255. ISBN 0-7923-3740-9.
  81. ^ Skoog, D.A; West, D.M.; Holler, J.F.; Crouch, S.R. (2004). Fundamentals of Analytical Chemistry (8th ed.). Thomson Brooks/Cole. ISBN 0-03-035523-0. Chapter 9-6: Acid Rain and the Buffer Capacity of Lakes
  82. ^ Stumm, W.; Morgan, J.J. (1996). Water Chemistry. New York: Wiley. ISBN 0-471-05196-9.
  83. ^ Snoeyink, V.L.; Jenkins, D. (1980). Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. New York: Wiley. ISBN 0-471-51185-4.
  84. ^ Millero, F.J. (2006). Chemical Oceanography (3rd ed.). London: Taylor and Francis. ISBN 0-8493-2280-4.
  85. ^ Millero, F.J.; Liu, X. (2002). "The Solubility of Iron in Seawater". Marine Chemistry. 77 (1): 43–54. Bibcode:2002MarCh..77...43L. doi:10.1016/S0304-4203(01)00074-3.
  86. ^ Speight, J.G. (2005). Lange's Handbook of Chemistry (18th ed.). McGraw–Hill. ISBN 0-07-143220-5. Chapter 8

Further reading edit

  • Albert, A.; Serjeant, E.P. (1971). The Determination of Ionization Constants: A Laboratory Manual. Chapman & Hall. ISBN 0-412-10300-1. (Previous edition published as Ionization constants of acids and bases. London (UK): Methuen. 1962.)
  • Atkins, P.W.; Jones, L. (2008). Chemical Principles: The Quest for Insight (4th ed.). W.H. Freeman. ISBN 978-1-4292-0965-6.
  • Housecroft, C. E.; Sharpe, A. G. (2008). Inorganic Chemistry (3rd ed.). Prentice Hall. ISBN 978-0-13-175553-6. (Non-aqueous solvents)
  • Hulanicki, A. (1987). Reactions of Acids and Bases in Analytical Chemistry. Horwood. ISBN 0-85312-330-6. (translation editor: Mary R. Masson)
  • Perrin, D.D.; Dempsey, B.; Serjeant, E.P. (1981). pKa Prediction for Organic Acids and Bases. Chapman & Hall. ISBN 0-412-22190-X.
  • Reichardt, C. (2003). Solvents and Solvent Effects in Organic Chemistry (3rd ed.). Wiley-VCH. ISBN 3-527-30618-8. Chapter 4: Solvent Effects on the Position of Homogeneous Chemical Equilibria.
  • Skoog, D.A.; West, D.M.; Holler, J.F.; Crouch, S.R. (2004). Fundamentals of Analytical Chemistry (8th ed.). Thomson Brooks/Cole. ISBN 0-03-035523-0.

External links edit

acid, dissociation, constant, acid, base, equilibrium, redirects, here, acid, base, balance, physiology, acid, base, homeostasis, redirects, here, other, uses, disambiguation, chemistry, acid, dissociation, constant, also, known, acidity, constant, acid, ioniz. Acid base equilibrium redirects here For acid base balance in physiology see Acid base homeostasis pKa redirects here For other uses see PKA disambiguation In chemistry an acid dissociation constant also known as acidity constant or acid ionization constant denoted K a displaystyle K a is a quantitative measure of the strength of an acid in solution It is the equilibrium constant for a chemical reaction HA A H displaystyle ce HA lt gt A H known as dissociation in the context of acid base reactions The chemical species HA is an acid that dissociates into A the conjugate base of the acid and a hydrogen ion H a The system is said to be in equilibrium when the concentrations of its components will not change over time because both forward and backward reactions are occurring at the same rate 1 The dissociation constant is defined by b K a A H H A displaystyle K text a mathrm frac A H HA or p K a log 10 K a log 10 HA A H displaystyle mathrm p K ce a log 10 K text a log 10 frac ce HA ce A ce H where quantities in square brackets represent the concentrations of the species at equilibrium c 2 As a simple example for a weak acid with Ka 10 5 log Ka is the exponent which is 5 so that pKa 5 And for acetic acid with Ka 1 8 x 10 5 pKa is close to 5 A higher Ka corresponds to a stronger acid which is more dissociated at equilibrium For the more convenient logarithmic scale a lower pKa means a stronger acid Contents 1 Theoretical background 2 Definitions 3 Equilibrium constant 3 1 Cumulative and stepwise constants 3 2 Association and dissociation constants 3 3 Temperature dependence 3 4 Dimensionality 4 Strong acids and bases 5 Monoprotic acids 6 Polyprotic acids 6 1 Isoelectric point 7 Bases and basicity 7 1 Basicity expressed as dissociation constant of conjugate acid 8 Amphoteric substances 8 1 Water self ionization 9 Acidity in nonaqueous solutions 9 1 Mixed solvents 10 Factors that affect pKa values 10 1 Thermodynamics 11 Experimental determination 11 1 Micro constants 12 Applications and significance 13 Values for common substances 14 See also 15 Notes 16 References 17 Further reading 18 External linksTheoretical background editThe acid dissociation constant for an acid is a direct consequence of the underlying thermodynamics of the dissociation reaction the pKa value is directly proportional to the standard Gibbs free energy change for the reaction The value of the pKa changes with temperature and can be understood qualitatively based on Le Chatelier s principle when the reaction is endothermic Ka increases and pKa decreases with increasing temperature the opposite is true for exothermic reactions The value of pKa also depends on molecular structure of the acid in many ways For example Pauling proposed two rules one for successive pKa of polyprotic acids see Polyprotic acids below and one to estimate the pKa of oxyacids based on the number of O and OH groups see Factors that affect pKa values below Other structural factors that influence the magnitude of the acid dissociation constant include inductive effects mesomeric effects and hydrogen bonding Hammett type equations have frequently been applied to the estimation of pKa 3 4 The quantitative behaviour of acids and bases in solution can be understood only if their pKa values are known In particular the pH of a solution can be predicted when the analytical concentration and pKa values of all acids and bases are known conversely it is possible to calculate the equilibrium concentration of the acids and bases in solution when the pH is known These calculations find application in many different areas of chemistry biology medicine and geology For example many compounds used for medication are weak acids or bases and a knowledge of the pKa values together with the octanol water partition coefficient can be used for estimating the extent to which the compound enters the blood stream Acid dissociation constants are also essential in aquatic chemistry and chemical oceanography where the acidity of water plays a fundamental role In living organisms acid base homeostasis and enzyme kinetics are dependent on the pKa values of the many acids and bases present in the cell and in the body In chemistry a knowledge of pKa values is necessary for the preparation of buffer solutions and is also a prerequisite for a quantitative understanding of the interaction between acids or bases and metal ions to form complexes Experimentally pKa values can be determined by potentiometric pH titration but for values of pKa less than about 2 or more than about 11 spectrophotometric or NMR measurements may be required due to practical difficulties with pH measurements Definitions editAccording to Arrhenius s original molecular definition an acid is a substance that dissociates in aqueous solution releasing the hydrogen ion H a proton 5 HA A H displaystyle ce HA lt gt A H nbsp The equilibrium constant for this dissociation reaction is known as a dissociation constant The liberated proton combines with a water molecule to give a hydronium or oxonium ion H3O naked protons do not exist in solution and so Arrhenius later proposed that the dissociation should be written as an acid base reaction HA H 2 O A H 3 O displaystyle ce HA H2O lt gt A H3O nbsp nbsp Acetic acid a weak acid donates a proton hydrogen ion highlighted in green to water in an equilibrium reaction to give the acetate ion and the hydronium ion Red oxygen black carbon white hydrogen Bronsted and Lowry generalised this further to a proton exchange reaction 6 7 8 acid base conjugate base conjugate acid displaystyle text acid text base ce lt gt text conjugate base text conjugate acid nbsp The acid loses a proton leaving a conjugate base the proton is transferred to the base creating a conjugate acid For aqueous solutions of an acid HA the base is water the conjugate base is A and the conjugate acid is the hydronium ion The Bronsted Lowry definition applies to other solvents such as dimethyl sulfoxide the solvent S acts as a base accepting a proton and forming the conjugate acid SH HA S A SH displaystyle ce HA S lt gt A SH nbsp In solution chemistry it is common to use H as an abbreviation for the solvated hydrogen ion regardless of the solvent In aqueous solution H denotes a solvated hydronium ion rather than a proton 9 10 The designation of an acid or base as conjugate depends on the context The conjugate acid BH of a base B dissociates according to BH OH B H 2 O displaystyle ce BH OH lt gt B H2O nbsp which is the reverse of the equilibrium H 2 O acid B base OH conjugate base BH conjugate acid displaystyle ce H2O text acid ce B text base ce lt gt OH text conjugate base ce BH text conjugate acid nbsp The hydroxide ion OH a well known base is here acting as the conjugate base of the acid water Acids and bases are thus regarded simply as donors and acceptors of protons respectively A broader definition of acid dissociation includes hydrolysis in which protons are produced by the splitting of water molecules For example boric acid B OH 3 produces H3O as if it were a proton donor 11 but it has been confirmed by Raman spectroscopy that this is due to the hydrolysis equilibrium 12 B OH 3 2 H 2 O B OH 4 H 3 O displaystyle ce B OH 3 2 H2O lt gt B OH 4 H3O nbsp Similarly metal ion hydrolysis causes ions such as Al H2O 6 3 to behave as weak acids 13 Al H 2 O 6 3 H 2 O Al H 2 O 5 OH 2 H 3 O displaystyle ce Al H2O 6 3 H2O lt gt Al H2O 5 OH 2 H3O nbsp According to Lewis s original definition an acid is a substance that accepts an electron pair to form a coordinate covalent bond 14 Equilibrium constant editAn acid dissociation constant is a particular example of an equilibrium constant The dissociation of a monoprotic acid HA in dilute solution can be written as HA A H displaystyle ce HA lt gt A H nbsp The thermodynamic equilibrium constant K displaystyle K ominus nbsp can be defined by 15 K A H HA displaystyle K ominus frac ce A ce H ce HA nbsp where X represents the activity at equilibrium of the chemical species X K displaystyle K ominus nbsp is dimensionless since activity is dimensionless Activities of the products of dissociation are placed in the numerator activities of the reactants are placed in the denominator See activity coefficient for a derivation of this expression nbsp Variation of pKa of acetic acid with ionic strength Since activity is the product of concentration and activity coefficient g the definition could also be written as K A H HA G G g A g H g HA displaystyle K ominus frac ce A ce H ce HA Gamma quad Gamma frac gamma ce A gamma ce H gamma ce HA nbsp where HA displaystyle text HA nbsp represents the concentration of HA and G displaystyle Gamma nbsp is a quotient of activity coefficients To avoid the complications involved in using activities dissociation constants are determined where possible in a medium of high ionic strength that is under conditions in which G displaystyle Gamma nbsp can be assumed to be always constant 15 For example the medium might be a solution of 0 1 molar M sodium nitrate or 3 M potassium perchlorate With this assumption K a K G A H H A displaystyle K text a frac K ominus Gamma mathrm frac A H HA nbsp p K a log 10 A H HA log 10 HA A H displaystyle mathrm p K ce a log 10 frac ce A ce H ce HA log 10 frac ce HA ce A ce H nbsp is obtained Note however that all published dissociation constant values refer to the specific ionic medium used in their determination and that different values are obtained with different conditions as shown for acetic acid in the illustration above When published constants refer to an ionic strength other than the one required for a particular application they may be adjusted by means of specific ion theory SIT and other theories 16 Cumulative and stepwise constants edit A cumulative equilibrium constant denoted by b displaystyle mathrm beta nbsp is related to the product of stepwise constants denoted by K displaystyle mathrm K nbsp For a dibasic acid the relationship between stepwise and overall constants is as follows H 2 A A 2 2 H displaystyle ce H2A lt gt A 2 2H nbsp b 2 H 2 A A 2 H 2 displaystyle beta 2 frac ce H2A ce A 2 ce H 2 nbsp log b 2 p K a 1 p K a 2 displaystyle log beta 2 mathrm p K ce a1 mathrm p K ce a2 nbsp Note that in the context of metal ligand complex formation the equilibrium constants for the formation of metal complexes are usually defined as association constants In that case the equilibrium constants for ligand protonation are also defined as association constants The numbering of association constants is the reverse of the numbering of dissociation constants in this example log b 1 p K a 2 displaystyle log beta 1 mathrm p K ce a2 nbsp Association and dissociation constants edit When discussing the properties of acids it is usual to specify equilibrium constants as acid dissociation constants denoted by Ka with numerical values given the symbol pKa K dissoc A H HA p K a log K dissoc displaystyle K text dissoc frac ce A H ce HA mathrm p K text a log K text dissoc nbsp On the other hand association constants are used for bases K assoc HA A H displaystyle K text assoc frac ce HA ce A H nbsp However general purpose computer programs that are used to derive equilibrium constant values from experimental data use association constants for both acids and bases Because stability constants for a metal ligand complex are always specified as association constants ligand protonation must also be specified as an association reaction 17 The definitions show that the value of an acid dissociation constant is the reciprocal of the value of the corresponding association constant K dissoc 1 K assoc displaystyle K text dissoc frac 1 K text assoc nbsp log K dissoc log K assoc displaystyle log K text dissoc log K text assoc nbsp p K dissoc p K assoc displaystyle mathrm p K text dissoc mathrm p K text assoc nbsp Notes For a given acid or base in water pKa pKb pKw the self ionization constant of water The association constant for the formation of a supramolecular complex may be denoted as Ka in such cases a stands for association not acid For polyprotic acids the numbering of stepwise association constants is the reverse of the numbering of the dissociation constants For example for phosphoric acid details in the polyprotic acids section below log K assoc 1 p K dissoc 3 log K assoc 2 p K dissoc 2 log K assoc 3 p K dissoc 1 displaystyle begin aligned log K text assoc 1 amp mathrm p K text dissoc 3 log K text assoc 2 amp mathrm p K text dissoc 2 log K text assoc 3 amp mathrm p K text dissoc 1 end aligned nbsp dd Temperature dependence edit All equilibrium constants vary with temperature according to the van t Hoff equation 18 d ln K d T D H R T 2 displaystyle frac mathrm d ln left K right mathrm d T frac Delta H ominus RT 2 nbsp R displaystyle R nbsp is the gas constant and T displaystyle T nbsp is the absolute temperature Thus for exothermic reactions the standard enthalpy change D H displaystyle Delta H ominus nbsp is negative and K decreases with temperature For endothermic reactions D H displaystyle Delta H ominus nbsp is positive and K increases with temperature The standard enthalpy change for a reaction is itself a function of temperature according to Kirchhoff s law of thermochemistry D H T p D C p displaystyle left frac partial Delta H partial T right p Delta C p nbsp where D C p displaystyle Delta C p nbsp is the heat capacity change at constant pressure In practice D H displaystyle Delta H ominus nbsp may be taken to be constant over a small temperature range Dimensionality edit In the equation K a A H H A displaystyle K mathrm a mathrm frac A H HA nbsp Ka appears to have dimensions of concentration However since D G R T ln K displaystyle Delta G RT ln K nbsp the equilibrium constant K displaystyle K nbsp cannot have a physical dimension This apparent paradox can be resolved in various ways Assume that the quotient of activity coefficients has a numerical value of 1 so that K displaystyle K nbsp has the same numerical value as the thermodynamic equilibrium constant K displaystyle K ominus nbsp Express each concentration value as the ratio c c0 where c0 is the concentration in a hypothetical standard state with a numerical value of 1 by definition 19 Express the concentrations on the mole fraction scale Since mole fraction has no dimension the quotient of concentrations will by definition be a pure number The procedures 1 and 2 give identical numerical values for an equilibrium constant Furthermore since a concentration c i displaystyle c i nbsp is simply proportional to mole fraction x i displaystyle x i nbsp and density r displaystyle rho nbsp c i x i r M displaystyle c i frac x i rho M nbsp and since the molar mass M displaystyle M nbsp is a constant in dilute solutions an equilibrium constant value determined using 3 will be simply proportional to the values obtained with 1 and 2 It is common practice in biochemistry to quote a value with a dimension as for example Ka 30 mM in order to indicate the scale millimolar mM or micromolar mM of the concentration values used for its calculation Strong acids and bases editAn acid is classified as strong when the concentration of its undissociated species is too low to be measured 6 Any aqueous acid with a pKa value of less than 0 is almost completely deprotonated and is considered a strong acid 20 All such acids transfer their protons to water and form the solvent cation species H3O in aqueous solution so that they all have essentially the same acidity a phenomenon known as solvent leveling 21 22 They are said to be fully dissociated in aqueous solution because the amount of undissociated acid in equilibrium with the dissociation products is below the detection limit Likewise any aqueous base with an association constant pKb less than about 0 corresponding to pKa greater than about 14 is leveled to OH and is considered a strong base 22 Nitric acid with a pK value of around 1 7 behaves as a strong acid in aqueous solutions with a pH greater than 1 23 At lower pH values it behaves as a weak acid pKa values for strong acids have been estimated by theoretical means 24 For example the pKa value of aqueous HCl has been estimated as 9 3 Monoprotic acids editSee also Acid Monoprotic acids nbsp Variation of the formation of a monoprotic acid AH and its conjugate base A with the difference between the pH and the pKa of the acid After rearranging the expression defining Ka and putting pH log10 H one obtains 25 p H p K a log A H A displaystyle mathrm pH mathrm p K text a log mathrm frac A HA nbsp This is the Henderson Hasselbalch equation from which the following conclusions can be drawn At half neutralization the ratio A HA 1 since log 1 0 the pH at half neutralization is numerically equal to pKa Conversely when pH pKa the concentration of HA is equal to the concentration of A The buffer region extends over the approximate range pKa 2 Buffering is weak outside the range pKa 1 At pH pKa 2 the substance is said to be fully protonated and at pH pKa 2 it is fully dissociated deprotonated If the pH is known the ratio may be calculated This ratio is independent of the analytical concentration of the acid In water measurable pKa values range from about 2 for a strong acid to about 12 for a very weak acid or strong base A buffer solution of a desired pH can be prepared as a mixture of a weak acid and its conjugate base In practice the mixture can be created by dissolving the acid in water and adding the requisite amount of strong acid or base When the pKa and analytical concentration of the acid are known the extent of dissociation and pH of a solution of a monoprotic acid can be easily calculated using an ICE table Polyprotic acids edit nbsp Phosphoric acid speciationA polyprotic acid is a compound which may lose more than 1 proton Stepwise dissociation constants are each defined for the loss of a single proton The constant for dissociation of the first proton may be denoted as Ka1 and the constants for dissociation of successive protons as Ka2 etc Phosphoric acid H3PO4 is an example of a polyprotic acid as it can lose three protons Equilibrium pK definition and value 26 H 3 PO 4 H 2 PO 4 H displaystyle ce H3PO4 lt gt H2PO4 H nbsp p K a 1 log 10 H 3 PO 4 H 2 PO 4 H 2 14 displaystyle mathrm p K ce a1 log 10 frac ce H 3PO 4 ce H 2PO 4 ce H 2 14 nbsp H 2 PO 4 HPO 4 2 H displaystyle ce H2PO4 lt gt HPO4 2 H nbsp p K a 2 log 10 H 2 PO 4 HPO 4 2 H 7 2 displaystyle mathrm p K ce a2 log 10 frac ce H 2PO 4 ce HPO 4 2 ce H 7 2 nbsp HPO 4 2 PO 4 3 H displaystyle ce HPO4 2 lt gt PO4 3 H nbsp p K a 3 log 10 HPO 4 2 PO 4 3 H 12 37 displaystyle mathrm p K ce a3 log 10 frac ce HPO4 2 ce PO4 3 ce H 12 37 nbsp When the difference between successive pK values is about four or more as in this example each species may be considered as an acid in its own right 27 In fact salts of H2 PO 4 may be crystallised from solution by adjustment of pH to about 5 5 and salts of HPO2 4 may be crystallised from solution by adjustment of pH to about 10 The species distribution diagram shows that the concentrations of the two ions are maximum at pH 5 5 and 10 nbsp species formation calculated with the program HySS for a 10 millimolar solution of citric acid pKa1 3 13 pKa2 4 76 pKa3 6 40 When the difference between successive pK values is less than about four there is overlap between the pH range of existence of the species in equilibrium The smaller the difference the more the overlap The case of citric acid is shown at the right solutions of citric acid are buffered over the whole range of pH 2 5 to 7 5 According to Pauling s first rule successive pK values of a given acid increase pKa2 gt pKa1 28 For oxyacids with more than one ionizable hydrogen on the same atom the pKa values often increase by about 5 units for each proton removed 29 30 as in the example of phosphoric acid above It can be seen in the table above that the second proton is removed from a negatively charged species Since the proton carries a positive charge extra work is needed to remove it which is why pKa2 is greater than pKa1 pKa3 is greater than pKa2 because there is further charge separation When an exception to Pauling s rule is found it indicates that a major change in structure is also occurring In the case of VO 2 aq the vanadium is octahedral 6 coordinate whereas vanadic acid is tetrahedral 4 coordinate This means that four particles are released with the first dissociation but only two particles are released with the other dissociations resulting in a much greater entropy contribution to the standard Gibbs free energy change for the first reaction than for the others Equilibrium pKa VO 2 H 2 O 4 H 3 VO 4 H 2 H 2 O displaystyle ce VO2 H2O 4 lt gt H3VO4 H 2H2O nbsp p K a 1 4 2 displaystyle mathrm p K a 1 4 2 nbsp H 3 VO 4 H 2 VO 4 H displaystyle ce H3VO4 lt gt H2VO4 H nbsp p K a 2 2 60 displaystyle mathrm p K a 2 2 60 nbsp H 2 VO 4 HVO 4 2 H displaystyle ce H2VO4 lt gt HVO4 2 H nbsp p K a 3 7 92 displaystyle mathrm p K a 3 7 92 nbsp HVO 4 2 VO 4 3 H displaystyle ce HVO4 2 lt gt VO4 3 H nbsp p K a 4 13 27 displaystyle mathrm p K a 4 13 27 nbsp Isoelectric point edit Main article isoelectric point For substances in solution the isoelectric point pI is defined as the pH at which the sum weighted by charge value of concentrations of positively charged species is equal to the weighted sum of concentrations of negatively charged species In the case that there is one species of each type the isoelectric point can be obtained directly from the pK values Take the example of glycine defined as AH There are two dissociation equilibria to consider AH 2 AH H AH H K 1 AH 2 displaystyle ce AH2 lt gt AH H qquad AH H mathit K 1 AH2 nbsp AH A H A H K 2 AH displaystyle ce AH lt gt A H qquad A H mathit K 2 AH nbsp Substitute the expression for AH from the second equation into the first equation A H 2 K 1 K 2 AH 2 displaystyle ce A H 2 mathit K 1 mathit K 2 AH2 nbsp At the isoelectric point the concentration of the positively charged species AH 2 is equal to the concentration of the negatively charged species A so H 2 K 1 K 2 displaystyle ce H 2 K 1 K 2 nbsp Therefore taking cologarithms the pH is given by p I p K 1 p K 2 2 displaystyle mathrm p I frac mathrm p K 1 mathrm p K 2 2 nbsp pI values for amino acids are listed at proteinogenic amino acid When more than two charged species are in equilibrium with each other a full speciation calculation may be needed Bases and basicity editThe equilibrium constant Kb for a base is usually defined as the association constant for protonation of the base B to form the conjugate acid HB B H 2 O HB OH displaystyle ce B H2O lt gt HB OH nbsp Using similar reasoning to that used before K b H B O H B p K b log 10 K b displaystyle begin aligned K text b amp mathrm frac HB OH B mathrm p K text b amp log 10 left K text b right end aligned nbsp Kb is related to Ka for the conjugate acid In water the concentration of the hydroxide ion OH is related to the concentration of the hydrogen ion by Kw H OH therefore O H K w H displaystyle mathrm OH frac K mathrm w mathrm H nbsp Substitution of the expression for OH into the expression for Kb gives K b H B K w B H K w K a displaystyle K text b frac mathrm HB K text w mathrm B H frac K text w K text a nbsp When Ka Kb and Kw are determined under the same conditions of temperature and ionic strength it follows taking cologarithms that pKb pKw pKa In aqueous solutions at 25 C pKw is 13 9965 31 so p K b 14 p K a displaystyle mathrm p K text b approx 14 mathrm p K text a nbsp with sufficient accuracy for most practical purposes In effect there is no need to define pKb separately from pKa 32 but it is done here as often only pKb values can be found in the older literature For an hydrolyzed metal ion Kb can also be defined as a stepwise dissociation constant M p OH q M p OH q 1 OH displaystyle mathrm M p ce OH q leftrightharpoons mathrm M p ce OH q 1 ce OH nbsp K b M p OH q 1 OH M p OH q displaystyle K mathrm b frac mathrm M p ce OH q 1 ce OH mathrm M p ce OH q nbsp This is the reciprocal of an association constant for formation of the complex Basicity expressed as dissociation constant of conjugate acid edit Because the relationship pKb pKw pKa holds only in aqueous solutions though analogous relationships apply for other amphoteric solvents subdisciplines of chemistry like organic chemistry that usually deal with nonaqueous solutions generally do not use pKb as a measure of basicity Instead the pKa of the conjugate acid denoted by pKaH is quoted when basicity needs to be quantified For base B and its conjugate acid BH in equilibrium this is defined as p K a H B p K a BH log 10 B H BH displaystyle mathrm p K mathrm aH mathrm B mathrm p K mathrm a ce BH log 10 Big frac ce B ce H ce BH Big nbsp A higher value for pKaH corresponds to a stronger base For example the values pKaH C5H5N 5 25 and pKaH CH3CH2 3N 10 75 indicate that CH3CH2 3N triethylamine is a stronger base than C5H5N pyridine Amphoteric substances editAn amphoteric substance is one that can act as an acid or as a base depending on pH Water below is amphoteric Another example of an amphoteric molecule is the bicarbonate ion HCO 3 that is the conjugate base of the carbonic acid molecule H2CO3 in the equilibrium H 2 CO 3 H 2 O HCO 3 H 3 O displaystyle ce H2CO3 H2O lt gt HCO3 H3O nbsp but also the conjugate acid of the carbonate ion CO2 3 in the reverse of the equilibrium HCO 3 OH CO 3 2 H 2 O displaystyle ce HCO3 OH lt gt CO3 2 H2O nbsp Carbonic acid equilibria are important for acid base homeostasis in the human body An amino acid is also amphoteric with the added complication that the neutral molecule is subject to an internal acid base equilibrium in which the basic amino group attracts and binds the proton from the acidic carboxyl group forming a zwitterion NH 2 CHRCO 2 H NH 3 CHRCO 2 displaystyle ce NH2CHRCO2H lt gt NH3 CHRCO2 nbsp At pH less than about 5 both the carboxylate group and the amino group are protonated As pH increases the acid dissociates according to NH 3 CHRCO 2 H NH 3 CHRCO 2 H displaystyle ce NH3 CHRCO2H lt gt NH3 CHRCO2 H nbsp At high pH a second dissociation may take place NH 3 CHRCO 2 NH 2 CHRCO 2 H displaystyle ce NH3 CHRCO2 lt gt NH2CHRCO2 H nbsp Thus the amino acid molecule is amphoteric because it may either be protonated or deprotonated Water self ionization edit Main article Self ionization of water The water molecule may either gain or lose a proton It is said to be amphiprotic The ionization equilibrium can be written H 2 O OH H displaystyle ce H2O lt gt OH H nbsp where in aqueous solution H denotes a solvated proton Often this is written as the hydronium ion H3O but this formula is not exact because in fact there is solvation by more than one water molecule and species such as H5O 2 H7O 3 and H9O 4 are also present 33 The equilibrium constant is given by K a H O H H 2 O displaystyle K text a mathrm frac H OH H 2 O nbsp With solutions in which the solute concentrations are not very high the concentration H2O can be assumed to be constant regardless of solute s this expression may then be replaced by K w H O H displaystyle K text w mathrm H mathrm OH nbsp The self ionization constant of water Kw is thus just a special case of an acid dissociation constant A logarithmic form analogous to pKa may also be defined p K w log 10 K w displaystyle mathrm p K text w log 10 left K text w right nbsp pKw values for pure water at various temperatures 34 T C 0 5 10 15 20 25 30 35 40 45 50pKw 14 943 14 734 14 535 14 346 14 167 13 997 13 830 13 680 13 535 13 396 13 262These data can be modelled by a parabola with p K w 14 94 0 04209 T 0 0001718 T 2 displaystyle mathrm p K mathrm w 14 94 0 04209 T 0 0001718 T 2 nbsp From this equation pKw 14 at 24 87 C At that temperature both hydrogen and hydroxide ions have a concentration of 10 7 M Acidity in nonaqueous solutions editA solvent will be more likely to promote ionization of a dissolved acidic molecule in the following circumstances 35 It is a protic solvent capable of forming hydrogen bonds It has a high donor number making it a strong Lewis base It has a high dielectric constant relative permittivity making it a good solvent for ionic species pKa values of organic compounds are often obtained using the aprotic solvents dimethyl sulfoxide DMSO 35 and acetonitrile ACN 36 Solvent properties at 25 C Solvent Donor number 35 Dielectric constant 35 Acetonitrile 14 37Dimethylsulfoxide 30 47Water 18 78DMSO is widely used as an alternative to water because it has a lower dielectric constant than water and is less polar and so dissolves non polar hydrophobic substances more easily It has a measurable pKa range of about 1 to 30 Acetonitrile is less basic than DMSO and so in general acids are weaker and bases are stronger in this solvent Some pKa values at 25 C for acetonitrile ACN 37 38 39 and dimethyl sulfoxide DMSO 40 are shown in the following tables Values for water are included for comparison pKa values of acids HA A H ACN DMSO Waterp Toluenesulfonic acid 8 5 0 9 Strong2 4 Dinitrophenol 16 66 5 1 3 9Benzoic acid 21 51 11 1 4 2Acetic acid 23 51 12 6 4 756Phenol 29 14 18 0 9 99BH B H ACN DMSO WaterPyrrolidine 19 56 10 8 11 4Triethylamine 18 82 9 0 10 72Proton sponge 18 62 7 5 12 1Pyridine 12 53 3 4 5 2Aniline 10 62 3 6 4 6Ionization of acids is less in an acidic solvent than in water For example hydrogen chloride is a weak acid when dissolved in acetic acid This is because acetic acid is a much weaker base than water HCl CH 3 CO 2 H Cl CH 3 C OH 2 displaystyle ce HCl CH3CO2H lt gt Cl CH3C OH 2 nbsp acid base conjugate base conjugate acid displaystyle text acid text base ce lt gt text conjugate base text conjugate acid nbsp Compare this reaction with what happens when acetic acid is dissolved in the more acidic solvent pure sulfuric acid 41 H 2 SO 4 CH 3 CO 2 H HSO 4 CH 3 C OH 2 displaystyle ce H2SO4 CH3CO2H lt gt HSO4 CH3C OH 2 nbsp nbsp Dimerization of a carboxylic acid The unlikely geminal diol species CH3C OH 2 is stable in these environments For aqueous solutions the pH scale is the most convenient acidity function 42 Other acidity functions have been proposed for non aqueous media the most notable being the Hammett acidity function H0 for superacid media and its modified version H for superbasic media 43 In aprotic solvents oligomers such as the well known acetic acid dimer may be formed by hydrogen bonding An acid may also form hydrogen bonds to its conjugate base This process known as homoconjugation has the effect of enhancing the acidity of acids lowering their effective pKa values by stabilizing the conjugate base Homoconjugation enhances the proton donating power of toluenesulfonic acid in acetonitrile solution by a factor of nearly 800 44 In aqueous solutions homoconjugation does not occur because water forms stronger hydrogen bonds to the conjugate base than does the acid Mixed solvents edit nbsp pKa of acetic acid in dioxane water mixtures Data at 25 C from Pine et al 45 When a compound has limited solubility in water it is common practice in the pharmaceutical industry for example to determine pKa values in a solvent mixture such as water dioxane or water methanol in which the compound is more soluble 46 In the example shown at the right the pKa value rises steeply with increasing percentage of dioxane as the dielectric constant of the mixture is decreasing A pKa value obtained in a mixed solvent cannot be used directly for aqueous solutions The reason for this is that when the solvent is in its standard state its activity is defined as one For example the standard state of water dioxane mixture with 9 1 mixing ratio is precisely that solvent mixture with no added solutes To obtain the pKa value for use with aqueous solutions it has to be extrapolated to zero co solvent concentration from values obtained from various co solvent mixtures These facts are obscured by the omission of the solvent from the expression that is normally used to define pKa but pKa values obtained in a given mixed solvent can be compared to each other giving relative acid strengths The same is true of pKa values obtained in a particular non aqueous solvent such a DMSO A universal solvent independent scale for acid dissociation constants has not been developed since there is no known way to compare the standard states of two different solvents Factors that affect pKa values editPauling s second rule is that the value of the first pKa for acids of the formula XOm OH n depends primarily on the number of oxo groups m and is approximately independent of the number of hydroxy groups n and also of the central atom X Approximate values of pKa are 8 for m 0 2 for m 1 3 for m 2 and lt 10 for m 3 28 Alternatively various numerical formulas have been proposed including pKa 8 5m known as Bell s rule 29 47 pKa 7 5m 30 48 or pKa 9 7m 29 The dependence on m correlates with the oxidation state of the central atom X the higher the oxidation state the stronger the oxyacid For example pKa for HClO is 7 2 for HClO2 is 2 0 for HClO3 is 1 and HClO4 is a strong acid pKa 0 7 The increased acidity on adding an oxo group is due to stabilization of the conjugate base by delocalization of its negative charge over an additional oxygen atom 47 This rule can help assign molecular structure for example phosphorous acid having molecular formula H3PO3 has a pKa near 2 which suggested that the structure is HPO OH 2 as later confirmed by NMR spectroscopy and not P OH 3 which would be expected to have a pKa near 8 48 nbsp pKa values for acetic chloroacetic dichloroacetic and trichloroacetic acids Inductive effects and mesomeric effects affect the pKa values A simple example is provided by the effect of replacing the hydrogen atoms in acetic acid by the more electronegative chlorine atom The electron withdrawing effect of the substituent makes ionisation easier so successive pKa values decrease in the series 4 7 2 8 1 4 and 0 7 when 0 1 2 or 3 chlorine atoms are present 49 The Hammett equation provides a general expression for the effect of substituents 50 log Ka log K0a rs Ka is the dissociation constant of a substituted compound K0a is the dissociation constant when the substituent is hydrogen r is a property of the unsubstituted compound and s has a particular value for each substituent A plot of log Ka against s is a straight line with intercept log K0a and slope r This is an example of a linear free energy relationship as log Ka is proportional to the standard free energy change Hammett originally 51 formulated the relationship with data from benzoic acid with different substituents in the ortho and para positions some numerical values are in Hammett equation This and other studies allowed substituents to be ordered according to their electron withdrawing or electron releasing power and to distinguish between inductive and mesomeric effects 52 53 Alcohols do not normally behave as acids in water but the presence of a double bond adjacent to the OH group can substantially decrease the pKa by the mechanism of keto enol tautomerism Ascorbic acid is an example of this effect The diketone 2 4 pentanedione acetylacetone is also a weak acid because of the keto enol equilibrium In aromatic compounds such as phenol which have an OH substituent conjugation with the aromatic ring as a whole greatly increases the stability of the deprotonated form nbsp Fumaric acid nbsp Maleic acidStructural effects can also be important The difference between fumaric acid and maleic acid is a classic example Fumaric acid is E 1 4 but 2 enedioic acid a trans isomer whereas maleic acid is the corresponding cis isomer i e Z 1 4 but 2 enedioic acid see cis trans isomerism Fumaric acid has pKa values of approximately 3 0 and 4 5 By contrast maleic acid has pKa values of approximately 1 5 and 6 5 The reason for this large difference is that when one proton is removed from the cis isomer maleic acid a strong intramolecular hydrogen bond is formed with the nearby remaining carboxyl group This favors the formation of the maleate H and it opposes the removal of the second proton from that species In the trans isomer the two carboxyl groups are always far apart so hydrogen bonding is not observed 54 nbsp Proton spongeProton sponge 1 8 bis dimethylamino naphthalene has a pKa value of 12 1 It is one of the strongest amine bases known The high basicity is attributed to the relief of strain upon protonation and strong internal hydrogen bonding 55 56 Effects of the solvent and solvation should be mentioned also in this section It turns out these influences are more subtle than that of a dielectric medium mentioned above For example the expected by electronic effects of methyl substituents and observed in gas phase order of basicity of methylamines Me3N gt Me2NH gt MeNH2 gt NH3 is changed by water to Me2NH gt MeNH2 gt Me3N gt NH3 Neutral methylamine molecules are hydrogen bonded to water molecules mainly through one acceptor N HOH interaction and only occasionally just one more donor bond NH OH2 Hence methylamines are stabilized to about the same extent by hydration regardless of the number of methyl groups In stark contrast corresponding methylammonium cations always utilize all the available protons for donor NH OH2 bonding Relative stabilization of methylammonium ions thus decreases with the number of methyl groups explaining the order of water basicity of methylamines 4 Thermodynamics edit An equilibrium constant is related to the standard Gibbs energy change for the reaction so for an acid dissociation constant D G R T ln K a 2 303 R T p K a displaystyle Delta G ominus RT ln K text a approx 2 303RT mathrm p K text a nbsp R is the gas constant and T is the absolute temperature Note that pKa log Ka and 2 303 ln 10 At 25 C DG in kJ mol 1 5 708 pKa 1 kJ mol 1 1000 joules per mole Free energy is made up of an enthalpy term and an entropy term 11 D G D H T D S displaystyle Delta G ominus Delta H ominus T Delta S ominus nbsp The standard enthalpy change can be determined by calorimetry or by using the van t Hoff equation though the calorimetric method is preferable When both the standard enthalpy change and acid dissociation constant have been determined the standard entropy change is easily calculated from the equation above In the following table the entropy terms are calculated from the experimental values of pKa and DH The data were critically selected and refer to 25 C and zero ionic strength in water 11 Acids Compound Equilibrium pKa DG kJ mol 1 d DH kJ mol 1 TDS kJ mol 1 e HA Acetic acid HA H A 4 756 27 147 0 41 27 56H2A GlycineH H2A HA H 2 351 13 420 4 00 9 419HA H A 9 78 55 825 44 20 11 6H2A Maleic acid H2A HA H 1 92 10 76 1 10 9 85HA H A2 6 27 35 79 3 60 39 4H3A Citric acid H3A H2A H 3 128 17 855 4 07 13 78H2A HA2 H 4 76 27 176 2 23 24 9HA2 A3 H 6 40 36 509 3 38 39 9H3A Boric acid H3A H2A H 9 237 52 725 13 80 38 92H3A Phosphoric acid H3A H2A H 2 148 12 261 8 00 20 26H2A HA2 H 7 20 41 087 3 60 37 5HA2 A3 H 12 35 80 49 16 00 54 49HA Hydrogen sulfate HA A2 H 1 99 11 36 22 40 33 74H2A Oxalic acid H2A HA H 1 27 7 27 3 90 11 15HA A2 H 4 266 24 351 7 00 31 35 The hydrogen ion does not exist as such in solution It combines with a solvent molecule when the solvent is water a hydronium ion is formed H H2O H3O This reaction is quantitative and hence can be ignored in the context of chemical equilibrium It is common practice to quote pK values rather than K values pK log10 K pKa is often referred to as an acid dissociation constant but this is strictly speaking incorrect as pKa is the cologarithm of the dissociation constant It is implicit in this definition that the quotient of activity coefficients g A H g A g H displaystyle frac gamma AH gamma A gamma H nbsp is a constant with a value of 1 under a given set of experimental conditions DG 2 303RTpKa Computed here from DH and DG values supplied in the citation using TDS DG DH Conjugate acids of bases Compound Equilibrium pKa DH kJ mol 1 TDS kJ mol 1 B Ammonia HB B H 9 245 51 95 0 8205B Methylamine HB B H 10 645 55 34 5 422B Triethylamine HB B H 10 72 43 13 18 06The first point to note is that when pKa is positive the standard free energy change for the dissociation reaction is also positive Second some reactions are exothermic and some are endothermic but when DH is negative TDS is the dominant factor which determines that DG is positive Last the entropy contribution is always unfavourable DS lt 0 in these reactions Ions in aqueous solution tend to orient the surrounding water molecules which orders the solution and decreases the entropy The contribution of an ion to the entropy is the partial molar entropy which is often negative especially for small or highly charged ions 57 The ionization of a neutral acid involves formation of two ions so that the entropy decreases DS lt 0 On the second ionization of the same acid there are now three ions and the anion has a charge so the entropy again decreases Note that the standard free energy change for the reaction is for the changes from the reactants in their standard states to the products in their standard states The free energy change at equilibrium is zero since the chemical potentials of reactants and products are equal at equilibrium Experimental determination edit nbsp A calculated titration curve of oxalic acid titrated with a solution of sodium hydroxideSee also Determination of equilibrium constants The experimental determination of pKa values is commonly performed by means of titrations in a medium of high ionic strength and at constant temperature 58 A typical procedure would be as follows A solution of the compound in the medium is acidified with a strong acid to the point where the compound is fully protonated The solution is then titrated with a strong base until all the protons have been removed At each point in the titration pH is measured using a glass electrode and a pH meter The equilibrium constants are found by fitting calculated pH values to the observed values using the method of least squares 59 The total volume of added strong base should be small compared to the initial volume of titrand solution in order to keep the ionic strength nearly constant This will ensure that pKa remains invariant during the titration A calculated titration curve for oxalic acid is shown at the right Oxalic acid has pKa values of 1 27 and 4 27 Therefore the buffer regions will be centered at about pH 1 3 and pH 4 3 The buffer regions carry the information necessary to get the pKa values as the concentrations of acid and conjugate base change along a buffer region Between the two buffer regions there is an end point or equivalence point at about pH 3 This end point is not sharp and is typical of a diprotic acid whose buffer regions overlap by a small amount pKa2 pKa1 is about three in this example If the difference in pK values were about two or less the end point would not be noticeable The second end point begins at about pH 6 3 and is sharp This indicates that all the protons have been removed When this is so the solution is not buffered and the pH rises steeply on addition of a small amount of strong base However the pH does not continue to rise indefinitely A new buffer region begins at about pH 11 pKw 3 which is where self ionization of water becomes important It is very difficult to measure pH values of less than two in aqueous solution with a glass electrode because the Nernst equation breaks down at such low pH values To determine pK values of less than about 2 or more than about 11 spectrophotometric 60 61 or NMR 62 63 measurements may be used instead of or combined with pH measurements When the glass electrode cannot be employed as with non aqueous solutions spectrophotometric methods are frequently used 38 These may involve absorbance or fluorescence measurements In both cases the measured quantity is assumed to be proportional to the sum of contributions from each photo active species with absorbance measurements the Beer Lambert law is assumed to apply Isothermal titration calorimetry ITC may be used to determine both a pK value and the corresponding standard enthalpy for acid dissociation 64 Software to perform the calculations is supplied by the instrument manufacturers for simple systems Aqueous solutions with normal water cannot be used for 1H NMR measurements but heavy water D2O must be used instead 13C NMR data however can be used with normal water and 1H NMR spectra can be used with non aqueous media The quantities measured with NMR are time averaged chemical shifts as proton exchange is fast on the NMR time scale Other chemical shifts such as those of 31P can be measured Micro constants edit nbsp CysteineFor some polyprotic acids dissociation or association occurs at more than one nonequivalent site 4 and the observed macroscopic equilibrium constant or macroconstant is a combination of microconstants involving distinct species When one reactant forms two products in parallel the macroconstant is a sum of two microconstants K K X K Y displaystyle K K X K Y nbsp This is true for example for the deprotonation of the amino acid cysteine which exists in solution as a neutral zwitterion HS CH2 CH NH 3 COO The two microconstants represent deprotonation either at sulphur or at nitrogen and the macroconstant sum here is the acid dissociation constant K a K a SH K a NH 3 displaystyle K mathrm a K mathrm a ce SH K mathrm a ce NH3 nbsp 65 nbsp SpermineSimilarly a base such as spermine has more than one site where protonation can occur For example monoprotonation can occur at a terminal NH2 group or at internal NH groups The Kb values for dissociation of spermine protonated at one or other of the sites are examples of micro constants They cannot be determined directly by means of pH absorbance fluorescence or NMR measurements a measured Kb value is the sum of the K values for the micro reactions K b K terminal K internal displaystyle K text b K text terminal K text internal nbsp Nevertheless the site of protonation is very important for biological function so mathematical methods have been developed for the determination of micro constants 66 When two reactants form a single product in parallel the macroconstant 1 K 1 K X 1 K Y displaystyle 1 K 1 K X 1 K Y nbsp 65 For example the abovementioned equilibrium for spermine may be considered in terms of Ka values of two tautomeric conjugate acids with macroconstant In this case 1 K a 1 K a terminal 1 K a internal displaystyle 1 K text a 1 K text a text terminal 1 K text a text internal nbsp This is equivalent to the preceding expression since K b displaystyle K mathrm b nbsp is proportional to 1 K a displaystyle 1 K mathrm a nbsp When a reactant undergoes two reactions in series the macroconstant for the combined reaction is the product of the microconstant for the two steps For example the abovementioned cysteine zwitterion can lose two protons one from sulphur and one from nitrogen and the overall macroconstant for losing two protons is the product of two dissociation constants K K a SH K a NH 3 displaystyle K K mathrm a ce SH K mathrm a ce NH3 nbsp 65 This can also be written in terms of logarithmic constants as p K p K a SH p K a NH 3 displaystyle mathrm p K mathrm p K mathrm a ce SH mathrm p K mathrm a ce NH3 nbsp Applications and significance editA knowledge of pKa values is important for the quantitative treatment of systems involving acid base equilibria in solution Many applications exist in biochemistry for example the pKa values of proteins and amino acid side chains are of major importance for the activity of enzymes and the stability of proteins 67 Protein pKa values cannot always be measured directly but may be calculated using theoretical methods Buffer solutions are used extensively to provide solutions at or near the physiological pH for the study of biochemical reactions 68 the design of these solutions depends on a knowledge of the pKa values of their components Important buffer solutions include MOPS which provides a solution with pH 7 2 and tricine which is used in gel electrophoresis 69 70 Buffering is an essential part of acid base physiology including acid base homeostasis 71 and is key to understanding disorders such as acid base disorder 72 73 74 The isoelectric point of a given molecule is a function of its pK values so different molecules have different isoelectric points This permits a technique called isoelectric focusing 75 which is used for separation of proteins by 2 D gel polyacrylamide gel electrophoresis Buffer solutions also play a key role in analytical chemistry They are used whenever there is a need to fix the pH of a solution at a particular value Compared with an aqueous solution the pH of a buffer solution is relatively insensitive to the addition of a small amount of strong acid or strong base The buffer capacity 76 of a simple buffer solution is largest when pH pKa In acid base extraction the efficiency of extraction of a compound into an organic phase such as an ether can be optimised by adjusting the pH of the aqueous phase using an appropriate buffer At the optimum pH the concentration of the electrically neutral species is maximised such a species is more soluble in organic solvents having a low dielectric constant than it is in water This technique is used for the purification of weak acids and bases 77 A pH indicator is a weak acid or weak base that changes colour in the transition pH range which is approximately pKa 1 The design of a universal indicator requires a mixture of indicators whose adjacent pKa values differ by about two so that their transition pH ranges just overlap In pharmacology ionization of a compound alters its physical behaviour and macro properties such as solubility and lipophilicity log p For example ionization of any compound will increase the solubility in water but decrease the lipophilicity This is exploited in drug development to increase the concentration of a compound in the blood by adjusting the pKa of an ionizable group 78 Knowledge of pKa values is important for the understanding of coordination complexes which are formed by the interaction of a metal ion Mm acting as a Lewis acid with a ligand L acting as a Lewis base However the ligand may also undergo protonation reactions so the formation of a complex in aqueous solution could be represented symbolically by the reaction M H 2 O n m LH M H 2 O n 1 L m 1 H 3 O displaystyle ce M H2O mathit n m ce LH lt gt ce M H2O n 1 ce L m 1 ce H3O nbsp To determine the equilibrium constant for this reaction in which the ligand loses a proton the pKa of the protonated ligand must be known In practice the ligand may be polyprotic for example EDTA4 can accept four protons in that case all pKa values must be known In addition the metal ion is subject to hydrolysis that is it behaves as a weak acid so the pK values for the hydrolysis reactions must also be known 79 Assessing the hazard associated with an acid or base may require a knowledge of pKa values 80 For example hydrogen cyanide is a very toxic gas because the cyanide ion inhibits the iron containing enzyme cytochrome c oxidase Hydrogen cyanide is a weak acid in aqueous solution with a pKa of about 9 In strongly alkaline solutions above pH 11 say it follows that sodium cyanide is fully dissociated so the hazard due to the hydrogen cyanide gas is much reduced An acidic solution on the other hand is very hazardous because all the cyanide is in its acid form Ingestion of cyanide by mouth is potentially fatal independently of pH because of the reaction with cytochrome c oxidase In environmental science acid base equilibria are important for lakes 81 and rivers 82 83 for example humic acids are important components of natural waters Another example occurs in chemical oceanography 84 in order to quantify the solubility of iron III in seawater at various salinities the pKa values for the formation of the iron III hydrolysis products Fe OH 2 Fe OH 2 and Fe OH 3 were determined along with the solubility product of iron hydroxide 85 Values for common substances editThere are multiple techniques to determine the pKa of a chemical leading to some discrepancies between different sources Well measured values are typically within 0 1 units of each other Data presented here were taken at 25 C in water 7 86 More values can be found in the Thermodynamics section above A table of pKa of carbon acids measured in DMSO can be found on the page on carbanions Chemical Equilibrium pKaBH Adenine BH B H 4 17BH 2 BH H 9 65H3A Arsenic acid H3A H2A H 2 22H2A HA2 H 6 98HA2 A3 H 11 53HA Benzoic acid HA H A 4 204HA Butyric acid HA H A 4 82H2A Chromic acid H2A HA H 0 98HA A2 H 6 5B Codeine BH B H 8 17HA Cresol HA H A 10 29HA Formic acid HA H A 3 751HA Hydrofluoric acid HA H A 3 17HA Hydrocyanic acid HA H A 9 21HA Hydrogen selenide HA H A 3 89HA Hydrogen peroxide 90 HA H A 11 7HA Lactic acid HA H A 3 86HA Propionic acid HA H A 4 87HA Phenol HA H A 9 99H2A L Ascorbic Acid H2A HA H 4 17HA A2 H 11 57See also editAcidosis Acids in wine tartaric malic and citric are the principal acids in wine Alkalosis Arterial blood gas Chemical equilibrium Conductivity electrolytic Grotthuss mechanism how protons are transferred between hydronium ions and water molecules accounting for the exceptionally high ionic mobility of the proton animation Hammett acidity function a measure of acidity that is used for very concentrated solutions of strong acids including superacids Ion transport number Ocean acidification dissolution of atmospheric carbon dioxide affects seawater pH The reaction depends on total inorganic carbon and on solubility equilibria with solid carbonates such as limestone and dolomite Law of dilution pCO2 pH Predominance diagram relates to equilibria involving polyoxyanions pKa values are needed to construct these diagrams Proton affinity a measure of basicity in the gas phase Stability constants of complexes formation of a complex can often be seen as a competition between proton and metal ion for a ligand which is the product of dissociation of an acid Notes editReferences edit Whitten Kenneth W Gailey Kenneth D Davis Raymond E 1992 General Chemistry 4th ed Saunders College Publishing p 660 ISBN 0 03 072373 6 Petrucci Ralph H Harwood William S Herring F Geoffrey 2002 General Chemistry 8th ed Prentice Hall pp 667 8 ISBN 0 13 014329 4 Perrin DD Dempsey B Serjeant EP 1981 Chapter 3 Methods of pKa Prediction pKa Prediction for Organic Acids and Bases secondary London Chapman amp Hall pp 21 26 doi 10 1007 978 94 009 5883 8 ISBN 978 0 412 22190 3 a b c Fraczkiewicz R 2013 In Silico Prediction of Ionization In Reedijk J ed Reference Module in Chemistry Molecular Sciences and Chemical Engineering secondary Reference Module in Chemistry Molecular Sciences and Chemical Engineering Online Vol 5 Amsterdam the Netherlands Elsevier doi 10 1016 B978 0 12 409547 2 02610 X ISBN 9780124095472 Miessler Gary L Tarr Donald A 1991 Inorganic Chemistry 2nd ed Prentice Hall ISBN 0 13 465659 8 Chapter 6 Acid Base and Donor Acceptor Chemistry a b Bell R P 1973 The Proton in Chemistry 2nd ed London Chapman amp Hall ISBN 0 8014 0803 2 Includes discussion of many organic Bronsted acids a b c Shriver D F Atkins P W 1999 Inorganic Chemistry 3rd ed Oxford Oxford University Press ISBN 0 19 850331 8 Chapter 5 Acids and Bases Housecroft C E Sharpe A G 2008 Inorganic Chemistry 3rd ed Prentice Hall ISBN 978 0 13 175553 6 Chapter 6 Acids Bases and Ions in Aqueous Solution Headrick J M Diken E G Walters R S Hammer N I Christie R A Cui J Myshakin E M Duncan M A Johnson M A Jordan K D 2005 Spectral Signatures of Hydrated Proton Vibrations in Water Clusters Science 308 5729 1765 69 Bibcode 2005Sci 308 1765H doi 10 1126 science 1113094 PMID 15961665 S2CID 40852810 Smiechowski M Stangret J 2006 Proton hydration in aqueous solution Fourier transform infrared studies of HDO spectra J Chem Phys 125 20 204508 204522 Bibcode 2006JChPh 125t4508S doi 10 1063 1 2374891 PMID 17144716 a b c Goldberg R Kishore N Lennen R 2002 Thermodynamic Quantities for the Ionization Reactions of Buffers PDF J Phys Chem Ref Data 31 2 231 370 Bibcode 2002JPCRD 31 231G doi 10 1063 1 1416902 Archived from the original PDF on 2008 10 06 Jolly William L 1984 Modern Inorganic Chemistry McGraw Hill pp 198 ISBN 978 0 07 032760 3 Burgess J 1978 Metal Ions in Solution Ellis Horwood ISBN 0 85312 027 7 Section 9 1 Acidity of Solvated Cations lists many pKa values Petrucci R H Harwood R S Herring F G 2002 General Chemistry 8th ed Prentice Hall ISBN 0 13 014329 4 p 698 a b Rossotti F J C Rossotti H 1961 The Determination of Stability Constants McGraw Hill Chapter 2 Activity and Concentration Quotients pp 5 10 Project Ionic Strength Corrections for Stability Constants International Union of Pure and Applied Chemistry Retrieved 2019 03 28 Rossotti Francis J C Rozotti Hazel 1961 The determination of stability constants and other equilibrium constants in solution New York McGraw Hill pp 5 10 ISBN 9781013909146 Archived from the original on 7 February 2020 Atkins P W de Paula J 2006 Physical Chemistry Oxford University Press ISBN 0 19 870072 5 Section 7 4 The Response of Equilibria to Temperature Petrucci Ralph H Harwood William S Herring F Geoffrey 2002 General chemistry principles and modern applications 8th ed Prentice Hall p 633 ISBN 0 13 014329 4 Are you wondering How using activities makes the equilibrium constant dimensionless Shriver D F Atkins P W 1999 Inorganic Chemistry 3rd ed Oxford University Press ISBN 0 19 850331 8 Sec 5 1c Strong and weak acids and bases Porterfield William W 1984 Inorganic Chemistry Addison Wesley p 260 ISBN 0 201 05660 7 a b Shriver D F Atkins P W 1999 Inorganic Chemistry 3rd ed Oxford University Press ISBN 0 19 850331 8 Sec 5 2 Solvent leveling Levanov A V Isaikina O Ya Lunin V V 2017 Dissociation constant of nitric acid Russian Journal of Physical Chemistry A 91 7 1221 1228 Bibcode 2017RJPCA 91 1221L doi 10 1134 S0036024417070196 S2CID 104093297 Trummal Aleksander Lipping Lauri Kaljurand Ivari Koppel Ilmar A Leito Ivo 2016 Acidity of Strong Acids in Water and Dimethyl Sulfoxide The Journal of Physical Chemistry A 120 20 3663 3669 Bibcode 2016JPCA 120 3663T doi 10 1021 acs jpca 6b02253 PMID 27115918 S2CID 29697201 Mehta Akul 22 October 2012 Henderson Hasselbalch Equation Derivation of pKa and pKb PharmaXChange Retrieved 16 November 2014 The values are for 25 C and 0 ionic strength Powell Kipton J Brown Paul L Byrne Robert H Gajda Tamas Hefter Glenn Sjoberg Staffan Wanner Hans 2005 Chemical speciation of environmentally significant heavy metals with inorganic ligands Part 1 The Hg2 Cl OH CO32 SO42 and PO43 aqueous systems Pure Appl Chem 77 4 739 800 doi 10 1351 pac200577040739 Brown T E Lemay H E Bursten B E Murphy C Woodward P 2008 Chemistry The Central Science 11th ed New York Prentice Hall p 689 ISBN 978 0 13 600617 6 a b Greenwood N N Earnshaw A 1997 Chemistry of the Elements 2nd ed Oxford Butterworth Heinemann p 50 ISBN 0 7506 3365 4 a b c Miessler Gary L Tarr Donald A 1999 Inorganic Chemistry 2nd ed Prentice Hall p 164 ISBN 0 13 465659 8 a b Huheey James E 1983 Inorganic Chemistry 3rd ed Harper amp Row p 297 ISBN 0 06 042987 9 Lide D R 2004 CRC Handbook of Chemistry and Physics Student Edition 84th ed CRC Press ISBN 0 8493 0597 7 Section D 152 Skoog Douglas A West Donald M Holler F James Crouch Stanley R 2014 Fundamentals of Analytical Chemistry 9th ed Brooks Cole p 212 ISBN 978 0 495 55828 6 Housecroft C E Sharpe A G 2004 Inorganic Chemistry 2nd ed Prentice Hall p 163 ISBN 978 0 13 039913 7 Harned H S Owen B B 1958 The Physical Chemistry of Electrolytic Solutions New York Reinhold Publishing Corp pp 634 649 752 754 a b c d Loudon G Marc 2005 Organic Chemistry 4th ed New York Oxford University Press pp 317 318 ISBN 0 19 511999 1 March J Smith M 2007 Advanced Organic Chemistry 6th ed New York John Wiley amp Sons ISBN 978 0 471 72091 1 Chapter 8 Acids and Bases Kutt A Movchun V Rodima T Dansauer T Rusanov E B Leito I Kaljurand I Koppel J Pihl V Koppel I Ovsjannikov G Toom L Mishima M Medebielle M Lork E Roschenthaler G V Koppel I A Kolomeitsev A A 2008 Pentakis trifluoromethyl phenyl a Sterically Crowded and Electron withdrawing Group Synthesis and Acidity of Pentakis trifluoromethyl benzene toluene phenol and aniline J Org Chem 73 7 2607 2620 doi 10 1021 jo702513w PMID 18324831 a b Kutt A Leito I Kaljurand I Soovali L Vlasov V M Yagupolskii L M Koppel I A 2006 A Comprehensive Self Consistent Spectrophotometric Acidity Scale of Neutral Bronsted Acids in Acetonitrile J Org Chem 71 7 2829 2838 doi 10 1021 jo060031y PMID 16555839 S2CID 8596886 Kaljurand I Kutt A Soovali L Rodima T Maemets V Leito I Koppel I A 2005 Extension of the Self Consistent Spectrophotometric Basicity Scale in Acetonitrile to a Full Span of 28 pKa Units Unification of Different Basicity Scales J Org Chem 70 3 1019 1028 doi 10 1021 jo048252w PMID 15675863 Bordwell pKa Table Acidity in DMSO Archived from the original on 9 October 2008 Retrieved 2008 11 02 Housecroft C E Sharpe A G 2008 Inorganic Chemistry 3rd ed Prentice Hall ISBN 978 0 13 175553 6 Chapter 8 Non Aqueous Media Rochester C H 1970 Acidity Functions Academic Press ISBN 0 12 590850 4 Olah G A Prakash S Sommer J 1985 Superacids New York Wiley Interscience ISBN 0 471 88469 3 Coetzee J F Padmanabhan G R 1965 Proton Acceptor Power and Homoconjugation of Mono and Diamines J Am Chem Soc 87 22 5005 5010 doi 10 1021 ja00950a006 Pine S H Hendrickson J B Cram D J Hammond G S 1980 Organic chemistry McGraw Hill p 203 ISBN 0 07 050115 7 Box K J Volgyi G Ruiz R Comer J E Takacs Novak K Bosch E Rafols C Roses M 2007 Physicochemical Properties of a New Multicomponent Cosolvent System for the pKa Determination of Poorly Soluble Pharmaceutical Compounds Helv Chim Acta 90 8 1538 1553 doi 10 1002 hlca 200790161 a b Housecroft Catherine E Sharpe Alan G 2005 Inorganic chemistry 2nd ed Harlow U K Pearson Prentice Hall pp 170 171 ISBN 0 13 039913 2 a b Douglas B McDaniel D H and Alexander J J Concepts and Models of Inorganic Chemistry 2nd ed Wiley 1983 p 526 ISBN 0 471 21984 3 Pauling L 1960 The nature of the chemical bond and the structure of molecules and crystals an introduction to modern structural chemistry 3rd ed Ithaca NY Cornell University Press p 277 ISBN 0 8014 0333 2 Pine S H Hendrickson J B Cram D J Hammond G S 1980 Organic Chemistry McGraw Hill ISBN 0 07 050115 7 Section 13 3 Quantitative Correlations of Substituent Effects Part B The Hammett Equation Hammett L P 1937 The Effect of Structure upon the Reactions of Organic Compounds Benzene Derivatives J Am Chem Soc 59 1 96 103 doi 10 1021 ja01280a022 Hansch C Leo A Taft R W 1991 A Survey of Hammett Substituent Constants and Resonance and Field Parameters Chem Rev 91 2 165 195 doi 10 1021 cr00002a004 S2CID 97583278 Shorter J 1997 Compilation and critical evaluation of structure reactivity parameters and equations Part 2 Extension of the Hammett s scale through data for the ionization of substituted benzoic acids in aqueous solvents at 25 C Technical Report Pure and Applied Chemistry 69 12 2497 2510 doi 10 1351 pac199769122497 S2CID 98814841 Pine S H Hendrickson J B Cram D J Hammond G S 1980 Organic chemistry McGraw Hill ISBN 0 07 050115 7 Section 6 2 Structural Effects on Acidity and Basicity Alder R W Bowman P S Steele W R S Winterman D R 1968 The Remarkable Basicity of 1 8 bis dimethylamino naphthalene Chem Commun 13 723 724 doi 10 1039 C19680000723 Alder R W 1989 Strain Effects on Amine Basicities Chem Rev 89 5 1215 1223 doi 10 1021 cr00095a015 Atkins Peter William De Paula Julio 2006 Atkins physical chemistry New York W H Freeman p 94 ISBN 978 0 7167 7433 4 Martell A E Motekaitis R J 1992 Determination and Use of Stability Constants Wiley ISBN 0 471 18817 4 Chapter 4 Experimental Procedure for Potentiometric pH Measurement of Metal Complex Equilibria Leggett D J 1985 Computational Methods for the Determination of Formation Constants Plenum ISBN 0 306 41957 2 Allen R I Box K J Comer J E A Peake C Tam K Y 1998 Multiwavelength Spectrophotometric Determination of Acid Dissociation Constants of Ionizable Drugs J Pharm Biomed Anal 17 4 5 699 712 doi 10 1016 S0731 7085 98 00010 7 PMID 9682153 Box K J Donkor R E Jupp P A Leader I P Trew D F Turner C H 2008 The Chemistry of Multi Protic Drugs Part 1 A Potentiometric Multi Wavelength UV and NMR pH Titrimetric Study of the Micro Speciation of SKI 606 J Pharm Biomed Anal 47 2 303 311 doi 10 1016 j jpba 2008 01 015 PMID 18314291 Popov K Ronkkomaki H Lajunen L H J 2006 Guidelines for NMR easurements for Determination of High and Low pKa Values PDF Pure Appl Chem 78 3 663 675 doi 10 1351 pac200678030663 S2CID 4823180 Szakacs Z Hagele G 2004 Accurate Determination of Low pK Values by 1H NMR Titration Talanta 62 4 819 825 doi 10 1016 j talanta 2003 10 007 PMID 18969368 Feig Andrew L ed 2016 Methods in Enzymology Calorimetry Elsevier 567 2 493 ISSN 0076 6879 a b c Splittgerber A G Chinander L L 1 February 1988 The spectrum of a dissociation intermediate of cysteine a biophysical chemistry experiment Journal of Chemical Education 65 2 167 Bibcode 1988JChEd 65 167S doi 10 1021 ed065p167 Frassineti C Alderighi L Gans P Sabatini A Vacca A Ghelli S 2003 Determination of Protonation Constants of Some Fluorinated Polyamines by Means of 13C NMR Data Processed by the New Computer Program HypNMR2000 Protonation Sequence in Polyamines Anal Bioanal Chem 376 7 1041 1052 doi 10 1007 s00216 003 2020 0 PMID 12845401 S2CID 14533024 Onufriev A Case D A Ullmann G M 2001 A Novel View of pH Titration in Biomolecules Biochemistry 40 12 3413 3419 doi 10 1021 bi002740q PMID 11297406 Good N E Winget G D Winter W Connolly T N Izawa S Singh R M M 1966 Hydrogen Ion Buffers for Biological Research Biochemistry 5 2 467 477 doi 10 1021 bi00866a011 PMID 5942950 Dunn M J 1993 Gel Electrophoresis Proteins Bios Scientific Publishers ISBN 1 872748 21 X Martin R 1996 Gel Electrophoresis Nucleic Acids Bios Scientific Publishers ISBN 1 872748 28 7 Brenner B M Stein J H eds 1979 Acid Base and Potassium Homeostasis Churchill Livingstone ISBN 0 443 08017 8 Scorpio R 2000 Fundamentals of Acids Bases Buffers amp Their Application to Biochemical Systems Kendall Hunt Pub Co ISBN 0 7872 7374 0 Beynon R J Easterby J S 1996 Buffer Solutions The Basics Oxford Oxford University Press ISBN 0 19 963442 4 Perrin D D Dempsey B 1974 Buffers for pH and Metal Ion Control London Chapman amp Hall ISBN 0 412 11700 2 Garfin D Ahuja S eds 2005 Handbook of Isoelectric Focusing and Proteomics Vol 7 Elsevier ISBN 0 12 088752 5 Hulanicki A 1987 Reactions of Acids and Bases in Analytical Chemistry Masson M R translation editor Horwood ISBN 0 85312 330 6 Eyal A M 1997 Acid Extraction by Acid Base Coupled Extractants Ion Exchange and Solvent Extraction A Series of Advances 13 31 94 Avdeef A 2003 Absorption and Drug Development Solubility Permeability and Charge State New York Wiley ISBN 0 471 42365 3 Beck M T Nagypal I 1990 Chemistry of Complex Equilibria Horwood ISBN 0 85312 143 5 van Leeuwen C J Hermens L M 1995 Risk Assessment of Chemicals An Introduction Springer pp 254 255 ISBN 0 7923 3740 9 Skoog D A West D M Holler J F Crouch S R 2004 Fundamentals of Analytical Chemistry 8th ed Thomson Brooks Cole ISBN 0 03 035523 0 Chapter 9 6 Acid Rain and the Buffer Capacity of Lakes Stumm W Morgan J J 1996 Water Chemistry New York Wiley ISBN 0 471 05196 9 Snoeyink V L Jenkins D 1980 Aquatic Chemistry Chemical Equilibria and Rates in Natural Waters New York Wiley ISBN 0 471 51185 4 Millero F J 2006 Chemical Oceanography 3rd ed London Taylor and Francis ISBN 0 8493 2280 4 Millero F J Liu X 2002 The Solubility of Iron in Seawater Marine Chemistry 77 1 43 54 Bibcode 2002MarCh 77 43L doi 10 1016 S0304 4203 01 00074 3 Speight J G 2005 Lange s Handbook of Chemistry 18th ed McGraw Hill ISBN 0 07 143220 5 Chapter 8Further reading editAlbert A Serjeant E P 1971 The Determination of Ionization Constants A Laboratory Manual Chapman amp Hall ISBN 0 412 10300 1 Previous edition published as Ionization constants of acids and bases London UK Methuen 1962 Atkins P W Jones L 2008 Chemical Principles The Quest for Insight 4th ed W H Freeman ISBN 978 1 4292 0965 6 Housecroft C E Sharpe A G 2008 Inorganic Chemistry 3rd ed Prentice Hall ISBN 978 0 13 175553 6 Non aqueous solvents Hulanicki A 1987 Reactions of Acids and Bases in Analytical Chemistry Horwood ISBN 0 85312 330 6 translation editor Mary R Masson Perrin D D Dempsey B Serjeant E P 1981 pKa Prediction for Organic Acids and Bases Chapman amp Hall ISBN 0 412 22190 X Reichardt C 2003 Solvents and Solvent Effects in Organic Chemistry 3rd ed Wiley VCH ISBN 3 527 30618 8 Chapter 4 Solvent Effects on the Position of Homogeneous Chemical Equilibria Skoog D A West D M Holler J F Crouch S R 2004 Fundamentals of Analytical Chemistry 8th ed Thomson Brooks Cole ISBN 0 03 035523 0 External links editAcidity Basicity Data in Nonaqueous Solvents Extensive bibliography of pKa values in DMSO acetonitrile THF heptane 1 2 dichloroethane and in the gas phase Curtipot All in one freeware for pH and acid base equilibrium calculations and for simulation and analysis of potentiometric titration curves with spreadsheets SPARC Physical Chemical property calculator Includes a database with aqueous non aqueous and gaseous phase pKa values than can be searched using SMILES or CAS registry numbers Aqueous Equilibrium Constants pKa values for various acid and bases Includes a table of some solubility products Free guide to pKa and log p interpretation and measurement Archived 2016 08 10 at the Wayback Machine Explanations of the relevance of these properties to pharmacology Free online prediction tool Marvin pKa log p log d etc From ChemAxon Chemicalize org List of predicted structure based properties pKa Chart 1 by David A Evans Retrieved from https en wikipedia org w index php title Acid dissociation constant amp oldid 1205360150, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.