„Savi disszociációs állandó” változatai közötti eltérés

A Wikipédiából, a szabad enciklopédiából
Tartalom törölve Tartalom hozzáadva
a új szócikk enwiki alapján
(Nincs különbség)

A lap 2009. augusztus 3., 17:27-kori változata

Az ecetsav - CH3COOH - egy metilcsoportból (CH3) és a hozzá kapcsolódó karboxilcsoportból (COOH) áll. A karboxilcsoport le tud adni egy protont egy vízmolekulának (H20), melynek során visszamarad egy acetát anion (CH3COO-) és keletkezik egy oxónium kation (H3O+). Ez egy egyensúlyi reakció, így a folyamat visszafelé is lejátszódhat.
A gyenge savak közé tartozó ecetsav egyensúlyi reakcióban protont (hidrogén ion, zölddel) ad át a víznek acetát ion és oxónium ion képződése közben. Piros: oxigén, fekete: szén, fehér: hidrogén.

A savi dissziciációs állandó, Ka, egy sav oldatbeli erősségének kvantitatív mértéke, a savak disszociációjának egyensúlyi állandója. Az egyensúlyt az alábbi módon írhatjuk fel:

HA ⇌ A + H+,

ahol HA egy általános sav, amelyből disszociáció során A - a sav konjugált bázisa -, valamint egy hidrogén ion vagy proton, H+ keletkezik, mely vizes oldatokban szolvatált oxónium ion formájában létezik.

Az ábrabeli példán a HA az ecetsav, A az acetát ion. Azt mondjuk, hogy a HA, A és H+ részecskék egyensúlyban vannak, ha a koncentrációjuk nem változik az idő múlásával. A disszociációs állandót rendszerint a [HA], [A] és [H+] jelölésű egyensúlyi koncentrációk hányadosaként írják fel:

Mivel Ka értéke több nagyságrendet átfoghat, ezért a gyakorlatban gyakran a savi disszociációs állandó logaritmusát használják. A −log10 Ka-val egyenlő pKa-t is szokták savi disszociációs állandónak nevezni:

Minél nagyobb pKa értéke, annál kisebb a disszociáció mértéke. A gyenge savak pKa értéke vízben körülbelül a −2 és 12 közötti tartományba esik. A körülbelül −2-nél kisebb pKa értékű savakat nevezzük erős savaknak. Az erős savak vizes oldatban csaknem teljesen disszociálnak: a nem disszociált sav kimutathatatlanná válik. Erős savak pKa értéke megbecsülhető elméleti úton, vagy extrapolálható olyan nem vizes oldószerekben végzett mérésekből, amelyekben a disszociációs állandó kisebb - ilyen oldószer például az acetonitril és a dimetil-szulfoxid.

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 Chatelier's principle: when the reaction is endothermic, the pKa decreases with increasing temperature; the opposite is true for exothermic reactions. The underlying structural factors that influence the magnitude of the acid dissociation constant include Pauling's rules for acidity constants, inductive effects, mesomeric effects, and hydrogen bonding.

The quantitative behaviour of acids and bases in solution can only be understood 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 water–octanol 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.

Definíciók

Arrhenius eredeti definíciója szerint a sav olyan anyag, mely vizes oldatban H+ hidrogénion (proton) leadása közben disszociál:[1]

HA ⇌ A + H+.

Ennek a disszociációs reakciónak az egyensúlyi állandója a disszociációs állandó.

A leadott proton egy vízmolekulával egyesülve H3O+ oxónium iont képez, ezért Arrhenius később javasolta, hogy a disszociációt sav-bázis reakcióként írják fel:

HA + H2O ⇌ A + H3O+.

A Brønsted–Lowry sav-bázis elmélet tovább általánosítva protoncsere reakciókról beszél:[2][3][4]

sav + bázis ⇌ konjugált bázis + konjugált sav.

A sav protont ad le és konjugált bázissá alakul, a protont felveszi a bázis, amely így konjugált savvá alakul. Egy HA sav vizes oldatában a bázis a víz, a konjugált bázis az A és a konjugált sav a oxónium ion. A Brønsted–Lowry definíció más oldószerekre is alkalmazható, például dimetil-szulfoxidra: az S oldószer, a bázis, protont vesz fel és konjugált savvá (SH+) alakul.

Az oldatkémiában - az oldószertől függetlenül - gyakran használják a H+ rövidítést a szolvatált hidrogén ion jelölésére.

Vizes oldatban a H+ nem protont, hanem szolvatált oxónium iont jelöl.[5][6]

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 + H2O

which is the reverse of the equilibrium

H2O (acid) + B (base) ⇌ OH (conjugate base) + BH+ (conjugate acid).

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.

Water is amphiprotic: it can react as an acid or a base. Another example of an amphiprotic molecule is thebicarbonate ion HCO3 which is the conjugate base of the carbonic acid molecule H2CO3 in the equilibrium

H2CO3 + H2O ⇌ HCO3 + H3O+

but also the conjugate acid of the carbonate ion CO32− in (the reverse of) the equilibrium

HCO3 + OH ⇌ CO32− + H2O.

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

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) acts as a weak acid, even though it is not a proton donor, because of the hydrolysis equilibrium

B(OH)3 + 2 H2O ⇌ B(OH)4 + H3O+.

Similarly, metal ion hydrolysis causes ions such as [Al(H2O)6]3+ to behave as weak acids:[7]

[Al(H2O)6]3+ +H2O ⇌ [Al(H2O)5(OH)]2+ + H3O+.

Egyensúlyi állandó

An acid dissociation constant is a particular example of an equilibrium constant. For the specific equilibrium between a monoprotic acid, HA and its conjugate base A, in water,

HA + H2O ⇌ A + H3O+

the thermodynamic equilibrium constant, K can be defined by[8]

where {A} is the activity of the chemical species A etc. K 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.

Illustration of the effect of ionic strength on the p K A of an acid. In this figure, the p K A of acetic acid decreases with increasing ionic strength, dropping from 4,8 in pure water (zero ionic strength) and becoming roughly constant at 4,45 for ionic strengths above 1 molar sodium nitrate, N A N O 3.
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 [HA] represents the concentration of HA and Γ 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 Γ can be assumed to be always constant.[8] For example, the medium might be a solution of 0,1 M sodium nitrate or 3 Mpotassium perchlorate (1 M = 1 mol·dm−3, a unit of molar concentration). Furthermore, in all but the most concentrated solutions it can be assumed that the concentration of water, [H2O], is constant, approximately 55 mol·dm−3. On dividing K by the constant terms and writing [H+] for the concentration of the hydronium ion the expression

is obtained. This is the definition in common use.[9] pKa is defined as −log10 Ka.

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.[10]

Although Ka appears to have the dimension of concentration it must in fact be dimensionless or it would not be possible to take its logarithm. The illusion is the result of omitting the constant term [H2O] from the defining expression. Nevertheless it is not unusual, particularly in texts relating to biochemical equilibria, to see a value quoted with a dimension as, for example, "Ka = 300 M".

Egyértékű savak

This figure plots the relative fractions of the protonated form A H of an acid to its deprotonated form, A minus, as the solution p H is varied about the value of the acid's p K A. When the p H equals the p K a, the amounts of the protonated and deprotonated forms are equal. When the p H is one unit higher than the p K A, the ratio of concentrations of protonated to deprotonated forms is 10 to 1. When the p H is two units higher that ratio is 100 to 1. Conversely, when the p H is one or two unit lower than the p K A, the ratio is 1 to ten or 1 to 100. The exact percentage of each form may be determined from the Henderson-Hasselbalch equation.
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

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

  • At half-neutralization [AH]/[A] = 1; since log(1) =0 , the pH at half-neutralization is numerically equal to pKa. Conversely, when pH = pKa, the concentration of AH is equal to the concentration of A.
  • The buffer region extends over the approximate range pKa ± 2, though buffering is weak outside the range pKa ± 1. At pKa ± 1, [AH]/[A] = 10 or 1/10.
  • If the pH is known, the ratio [AH]:[A] 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). All acids with a pKa value of less than −2 are more than 99% dissociated at pH 0 (1 M acid). This is known as solvent leveling since all such acids are brought to the same level of being strong acids, regardless of their pKa values. Likewise, all bases with a pKa value larger than the upper limit are more than 99% de-protonated at all attainable pH values and are classified as strong bases.[3]

An example of a strong acid is hydrochloric acid, HCl, which has a pKa value, estimated from thermodynamic quantities, of −9,3 in water.[11] The concentration of undissociated acid in a 1 mol·dm−3 solution will be less than 0.01% of the concentrations of the products of dissociation. Hydrochloric acid is said to be "fully dissociated" in aqueous solution because the amount of undissociated acid is imperceptible. 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.

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. The pKa of the acid must be less than two units different from the target pH.

Többértékű savak

Acids with more than one ionizable hydrogen atoms are called polyprotic acids, and have multiple deprotonation states, also called species. This image plots the relative percentages of the different protonation species of phosphoric acid H 3 P O 4 as a function of solution p H. Phosphoric acid has three ionizable hydrogen atoms whose p K A's are roughly 2, 7 and 12. Below p H 2, the triply protonated species H 3 P O 4 predominates; the double protonated species H 2 P O 4 minus predominates near p H 5; the singly protonated species H P O 4 2 minus predominates near p H 9 and the unprotonated species P O 4 3 minus predominates above p H 12.
% species' formation as a function of pH
This image plots the relative percentages of the protonation species of citric acid as a function of p H. Citric acid has three ionizable hydrogen atoms and thus three p K A values. Below the lowest p K A, the triply protonated species prevails; between the lowest and middle p K A, the doubly protonated form prevails; between the middle and highest p K A, the singly protonated form prevails; and above the highest p K A, the unprotonated form of citric acid is predominant.
% species formation calculated with the program HySS for a 10 millimolar solution of citric acid. pKa1=3,13, pKa2 = 4,76, pKa3=6,40.

Polyprotic acids are acids that can lose more than one proton. The constant for dissociation of the first proton may be denoted asKa1 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 pKa value
H3PO4 ⇌ H2PO4 + H+ pKa1 = 2,15
H2PO4 ⇌ HPO42− + H+ pKa2 = 7,20
HPO42− ⇌ PO43− + H+ pKa3 = 12,37

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;[12] In fact salts of H2PO4 may be crystallised from solution by adjustment of pH to about 5,5 and salts of HPO42− 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.

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.

It is generally true that successive pK values increase (Pauling's first rule).[13] For example, for a diprotic acid, H2A, the two equilibria are

H2A ⇌ HA + H+
HA ⇌ A2− + H+

it can be seen 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; that is the cause of the trend noted above. Phosphoric acid values (above) illustrate this rule, as do the values for vanadic acid, H3VO4. When an exception to the rule is found it indicates that a major change in structure is occurring. In the case of VO2+ (aq), the vanadium is octahedral, 6-coordinate, whereas vanadic acid is tetrahedral, 4-coordinate. This is the basis for an explanation of why pKa1 > pKa2 for vanadium(V) oxoacids.

equilibrium pKa value
[VO2(H2O)4]+ ⇌ H3VO4 + H+ + 2H2O pKa1 = 4,2
H3VO4 ⇌ H2VO4 + H+ pKa2 = 2,60
H2VO4 ⇌ HVO42− + H+ pKa3 = 7,92
HVO42− ⇌ VO43− + H+ pKa4 = 13,27

Water self-ionization

Water has both acidic and basic properties. The equilibrium constant for the equilibrium

2 H2O ⇌ OH + H3O+

is given by

When, as is usually the case, the concentration of water can be assumed to be constant, this expression may be replaced by

The value of Kw at SATP is 1,0··10−14. Theself-ionization constant of water, Kw, is thus just a special case of an acid dissociation constant.

Bases

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

B + H2O ⇌ HB+ + OH

Using similar reasoning to that used before


In water, the concentration of the hydroxide ion, [OH], is related to the concentration of the hydrogen ion byKw = [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,[14] so pKb ~ 14 − pKa.

In effect there is no need to define pKb separately from pKa, but it is done here because pKb values can be found in the older literature.

Hőmérsékletfüggés

All equilibrium constants vary with temperature according to the van 't Hoff equation[15]

R is the gas constant and T is the temperature in Kelvin. Thus, for exothermic reactions, (the standardenthalpy change, ΔH, is negative) K decreases with temperature, but for endothermicreactionsH is positive) K increases with temperature.

Acidity in nonaqueous solutions

A solvent will be more likely to promote ionization of a dissolved acidic molecule in the following circumstances.[16]

  1. It is a protic solvent, capable of forming hydrogen bonds.
  1. It has a high donor number, making it a strong Lewis base.
  1. 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)[16] and acetonitrile (AN).[17]

Oldószerek tulajdonságai 25°C-on
Oldószer Donorszám[16] Dielektromos állandó[16]
Acetonitril 14 37
Dimetil-szulfoxid 30 47
Víz 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 acids are generally weaker and bases are generally stronger in this solvent. Some pKa values at 25oC for acetonitrile (AN)[18][19][20] and dimethyl sulfoxide (DMSO)[21] are shown in the following tables. Values for water are included for comparison.

pKa values of acids
HA ⇌ A + H+ AN DMSO water
p-Toluenesulfonic acid 8,5 0,9 erős
2,4-Dinitrofenol 16,66 5,1 3,9
Benzoesav 21,51 11,1 4,2
Ecetsav 23,51 12,6 4,756
Fenol 29,14 18,0 9,99
BH+ ⇌ B + H+
Pyrrolidine 19,56 10,8 11,4
Trietilamin 18,82 9,0 10,72
Proton sponge            18,62 7,5 12,1
Piridin 12,53 3,4 5,2
Anilin 10,62 3,6 9,4


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

HCl + CH3CO2H ⇌ Cl + CH3C(OH)2+
acid + base ⇌ conjugate base + conjugate acid

Compare this reaction with what happens when acetic acid is dissolved in the more acidic solvent pure sulfuric acid[22]

H2SO4 + CH3CO2H ⇌ HSO4 + CH3C(OH)2+

The apparently unlikely geminal diol species CH3C(OH)2+ is stable in these environments. For aqueous solutions the pH scale is the most convenient acidity function.[23] Other acidity functions have been proposed for non-aqueous media, most notably the Hammett acidity function,H0, for superacid media and its modified version H for superbasic media.[24]

This image illustrates how two carboxylic acids, C O O H, can associate through mutual hydrogen bonds. The hydroxyl portion O H of each molecule forms a hydrogen bond to the carbonyl portion C O of the other.
Dimerization of a carboxylic acid

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

Mixed solvents

The p K A of acetic acid in the mixed solvent dioxane/water. p K A increases as the proportion of dioxane increases, primarily because the dielectric constant of the mixture decreases with increasing doxane content. A lower dielectric constant disfavors the dissociation of the uncharged acid into the charged ions, H + and C H 3 C O O minus, shifting the equilibrium to favor the uncharged protonated form C H 3 C O O H. Since the protonated form is the reactant not the product of the dissociation, this shift decreases the equilibrium constant K A, and increases P K A, its negative logarithm.
pKa of acetic acid in dioxane/water mixtures. Data at 25oC from Pine et al.[26]

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.[27] 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 9:1 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 which 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.

As of 2008, 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 which affect pKa values

Pauling's second rule states that the value of the first pKa for acids of the formula XOm(OH)n is approximately independent of n and X and is approximately 8 for m = 0, 2 for m = 1, −3 for m = 2 and < −10 for m = 3.[13] This 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.

Fumaric acid consists of two double-bonded carbon atoms capped on both sides by carboxylic acid groups C O O H; thus, its chemical formula is C O O H C H C H C O O H. The molecule has two ionizable hydrogen atoms and thus two p K As. The central double bond is in the trans configuration, which holds the two carboxylate groups apart. This contrasts with the cis isomer, maleic acid.
Fumaric acid
Maleic acid consists of two double-bonded carbon atoms capped on both sides by carboxylic acid groups C O O H; thus, its chemical formula is C O O H C H C H C O O H. It has two ionizable hydrogen atoms and thus two p K As. The central double bond is in the cis configuration. This holds the two carboxylate groups close enough so that when one group is protonated and the other deprotonated, a strong hydrogen bond can be formed between the two groups. This makes the mono-protonated species much more stable than the corresponding species of the trans isomer, fumaric acid.
Maleic acid
Proton sponge is a derivative of naphthalene with dimethylamino groups in the one and ten positions. This brings the two dimethyl amino groups into close proximity to each other.
proton sponge

With organic 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, 1,3 and 0,7 when 0,1, 2 or 3 chlorine atoms are present.[28] The Hammett equation, provides a general expression for the effect of substituents.[29]

log Ka = log Ka0 + ρσ.

Ka is the dissociation constant of a substituted compound, Ka0 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 logKa0 and slope ρ. This is an example of a linear free energy relationship as log Kais proportional to the standard fee energy change. Hammett originally[30] formulated the relationship with data from benzoic acid with different substiuents 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 orelectron-releasing power, and to distinguish between inductive and mesomeric effects.[31][32]

Alcohols do not normally behave as acids in water, but the presence of an 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 3,5 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.[33]

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.[34][35]

Termodinamika

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

ΔG = 2,303 RT pKa.

R is the gas constant and T is the temperature in Kelvin. 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 enthalpyterm and an entropy term.[36]

ΔG = ΔHTΔS

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.[36]

Acids
Compound Equilibrium pKa ΔH /kJ·mol−1 TΔS /kJ·mol−1
HA = Ecetsav HA ⇌ H+ + A 4,756 −0,41 27,56
H2A+ = GlycineH+ H2A+ ⇌ HA + H+ 2,351 4,00 9,419
HA ⇌ H+ + A 9,78 44,20 11,6
H2A = Maleinsav H2A ⇌ HA + H+ 1,92 1,10 9,85
HA ⇌ H+ + A2− 6,27 −3,60 39,4
H3A = Citromsav H3A ⇌ H2A + H+ 3,128 4,07 13,78
H2A ⇌ HA2− + H+ 4,76 2,23 24,9
HA2− ⇌ A3− + H+ 6,40 −3,38 39,9
HA = Bórsav HA ⇌ H+ + A 9,237 13,80 38,92
H3A = Foszforsav H3A ⇌ H2A + H+ 2,148 −8,00 20,26
H2A ⇌ HA2− + H+ 7,20 3,60 37,5
HA2− ⇌ A3− + H+ 12,35 16,00 54,49
HA = Hydrogen sulfate HA ⇌ A2− + H+ 1,99 −22,40 33,74
H2A = Oxálsav H2A ⇌ HA + H+ 1,27 −3,90 11,15
HA ⇌ A2− + H+ 4,266 7,00 31,35
Conjugate acid of bases
Compound Equilibrium pKa ΔH /kJ·mol−1 TΔS /kJ·mol−1
B = Ammónia HB+ ⇌ B + H+ 9,245 51,95 0,8205
B = Metilamin HB+ ⇌ B + H+ 10,645 55,34 5,422
B = Trietilamin 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, that is, dissociation of a weak acid is not a spontaneous process. Secondly some reactions are exothermicand some are endothermic, but when ΔH is negative−TΔS is the dominant factor which determines that ΔG is positive. Lastly, the entropy contribution is always unfavourable in these reactions.

Note that the standard free energy change for the reaction is for the changes from the reactants in their standard statesto 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

The image shows the titration curve of oxalic acid, showing the pH of the solution as a function of added base. There is a small inflection point at about pH 3 and then a large jump from pH 5 to pH 11, followed by another region of slowly increasing pH.
A calculated titration curve of oxalic acid titrated with a solution ofsodium 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.[37] 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.[38]

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, where the pH rises by about two units. 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[39] or NMR[9][40] measurements may be used instead of, or combined with, pH measurements.[41]

When the glass electrode cannot be employed, as with non-aqueous solutions, spectrophotometric methods are frequently used.[19] 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 lawis assumed to apply.

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

Spermine is a long, symmetrical molecule capped at both ends with amino groups N H 2. It has two N H groups symmetrically placed within the molecule, separated from each other by four methylene groups C H 2, and from the amino ends by three methylene groups. Thus, the full molecular formula is N H 2 C H 2 C H 2 C H 2 N H C H 2 C H 2 C H 2 C H 2 N H C H 2 C H 2 C H 2 N H 2.
spermine

A base such as spermine has a few different sites where protonation can occur. In this example the first proton can go on the terminal -NH2 group, or either of the internal -NH- groups. The pKa 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. Nevertheless, the site of protonation is very important for biological function, so mathematical methods have been developed for the determination of micro-constants.[42]

Applications and significance

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 acidside chains are of major importance for the activity of enzymes and the stability of proteins.[43] 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;[44] 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.[45][46] Buffering is an essential part of acid base physiology including acid-base homeostasis,[47] and is key to understanding disorders such as acid-base imbalance.[48][49][50] 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 focussing,[51] 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[52] 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.[53]

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 pKavalues 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.[54]

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(H2O)n]m+ +LH ⇌ [M(H2O)n−1L](m−1)+ + H3O+

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 tohydrolysis, that is, it behaves as a weak acid, so the pK values for the hydrolysis reactions must also be known.[55]

Assessing the hazard associated with an acid or base may require a knowledge of pKa values.[56] 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[57] and rivers;[58][59] for example, humic acids are important components of natural waters. Another example occurs in chemical oceanography:[60]

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.[61]

Values for common substances

Vegyületek pKa-jának meghatározására többféle módszer is létezik, ezért a különböző források között van némi eltérés.

A helyesen meghatározott értékek közötti eltérés jellemzően nem haladja meg a 0,1 egységet. Az alábbi adatok 25 °C-os vízre vonatkoznak.[3][62]

More values can be found in thermodynamics, above.

Chemical Name Equilibrium pKa
B = Adenin BH22+ ⇌ BH+ + H+ 4,17
BH+ ⇌ B + H+ 9,65
H3A = Arzénsav H3A ⇌ H2A + H+ 2,22
H2A ⇌ HA2− + H+ 6,98
HA2− ⇌ A3− + H+ 11,53
HA = Benzoesav HA ⇌ H+ + A 4,204
HA = Butánsav HA ⇌ H+ + A 4,82
H2A = Krómsav H2A ⇌ HA + H+ 0,98
HA ⇌ A2− + H+ 6,5
B = Kodein BH+ ⇌ B + H+ 8,17
HA = Cresol HA ⇌ H+ + A 10,29
HA = Hangyasav HA ⇌ H+ + A 3,751
HA = Hydrofluoric acid HA ⇌ H+ + A 3,17
HA = Hydrocyanic acid HA ⇌ H+ + A 9,21
HA = Hidrogén-szelenid HA ⇌ H+ + A 3,89
HA = Hidrogén-peroxid (90%) HA ⇌ H+ + A 11,7
HA = Tejsav HA ⇌ H+ + A 3,86
HA = Propánsav HA ⇌ H+ + A 4,87
HA = Fenol HA ⇌ H+ + A 9,99
H2A = L-(+)-Aszkorbinsav H2A ⇌ HA + H+ 4,17
HA ⇌ A2− + H+ 11,57

Lásd még

Hivatkozások

Ez a szócikk részben vagy egészben a Acid dissociation constant című angol Wikipédia-szócikk ezen változatának fordításán alapul. Az eredeti cikk szerkesztőit annak laptörténete sorolja fel. Ez a jelzés csupán a megfogalmazás eredetét és a szerzői jogokat jelzi, nem szolgál a cikkben szereplő információk forrásmegjelöléseként.

  1. Miessler, G.. Inorganic Chemistry, 2nd, Prentice Hall (1991). ISBN 0134656598  Chapter 6: Acid-Base and Donor-Acceptor Chemistry
  2. Bell, R.P.. The Proton in Chemistry, 2nd, London: Chapman & Hall (1973)  Includes discussion of many organic Brønsted acids.
  3. a b c Shriver, D.F, Atkins, P.W.. Inorganic Chemistry, 3rd, Oxford: Oxford University Press (1999). ISBN 0198503318  Chapter 5: Acids and Bases
  4. Housecroft, C.E., Sharpe, A.G.. Inorganic Chemistry, 3rd, Prentice Hall (2008). ISBN 0131755536  Chapter 6: Acids, Bases and Ions in Aqueous Solution
  5. 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, 1765–69. o. DOI:10.1126/science.1113094.  
  6. Smiechowski, M., Stangret, J. (2006). „Proton hydration in aqueous solution: Fourier transform infrared studies of HDO spectra”. J. Chem. Phys. 125, 204508–204522. o. DOI:10.1063/1.2374891.  
  7. Burgess, J.. Metal Ions in Solution. Ellis Horwood (1978). ISBN 0853120277  Section 9.1 "Acidity of Solvated Cations" lists many pKa values.
  8. a b Rossotti, F.J.C., Rossotti, H.. The Determination of Stability Constants. McGraw–Hill (1961)  Chapter 2: Activity and Concentration Quotients
  9. a b Popov, K., Ronkkomaki, H.; Lajunen, L.H.J. (2006). „Guidelines for NMR Measurements for Determination of High and Low pKa Values” (PDF). Pure Appl. Chem. 78 (3), 663–675. o. DOI:10.1351/pac200678030663.  
  10. Project: Ionic Strength Corrections for Stability Constants. International Union of Pure and Applied Chemistry. (Hozzáférés: 2008. november 23.)
  11. Dasent, W.E.. Inorganic Energetics: An Introduction. Cambridge University Press (1982). ISBN 0521284066  Chapter 5
  12. Brown, T.E., Lemay, H.E.; Bursten,B.E.; Murphy, C.; Woodward, P.. Chemistry: The Central Science, 11th, New York: Prentice-Hall, 689. o. (2008). ISBN 0136006175 
  13. a b Greenwood, N.N.. Az elemek kémiája, 1., Budapest: Nemzeti Tankönyvkiadó, 67. o. (1999). ISBN 963-18-9144-5 
  14. Lide, D.R.. CRC Handbook of Chemistry and Physics, Student Edition, 84th, CRC Press (2004). ISBN 0849305977  Section D–152
  15. Atkins, P.W., de Paula, J.. Physical Chemistry. Nemzeti Tankönyvkiadó (1998). ISBN 963-18-9163-1  9.4 fejezet: A hőmérséklet hatása az egyensúlyra
  16. a b c d Sablon:Loudon p. 317–318
  17. March, J., Smith, M.. Advanced Organic Chemistry, 6th, New York: John Wiley & Sons (2007). ISBN 978-0-471-72091-1  Chapter 8: Acids and Bases
  18. 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. o. DOI:10.1021/jo702513w.  
  19. 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. o. DOI:10.1021/jo060031y.  
  20. 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. o. DOI:10.1021/jo048252w.  
  21. Bordwell pKa Table (Acidity in DMSO). (Hozzáférés: 2008. november 2.)
  22. Housecroft, C.E., Sharpe, A.G.. Inorganic Chemistry, 3rd, Prentice Hall (2008). ISBN 0131755536  Chapter 8: Non-Aqueous Media
  23. Rochester, C.H.. Acidity Functions. Academic Press (1970). ISBN 0125908504 
  24. Olah, G.A, Prakash, S; Sommer, J. Superacids. New York: Wiley Interscience (1985). ISBN 0471884693 
  25. Coetzee, J.F., Padmanabhan, G.R. (1965). „Proton Acceptor Power and Homoconjugation of Mono- and Diamines”. J. Amer. Chem. Soc. 87, 5005–5010. o. DOI:10.1021/ja00950a006.  
  26. Pine, S.H., Hendrickson, J.B.; Cram, D.J.; Hammond, G.S. (1980). „Organic chemistry”, 203. o, Kiadó: McGraw–Hill.  
  27. 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. o. DOI:10.1002/hlca.200790161.  
  28. Pauling, L.. The nature of the chemical bond and the structure of molecules and crystals; an introduction to modern structural chemistry, 3rd, Ithaca (NY): Cornell University Press, 277. o. (1960) 
  29. Pine, S.H., Hendrickson, J.B.; Cram, D.J.; Hammond, G.S.. Organic Chemistry. McGraw–Hill (1980). ISBN 0070501157  Section 13-3: Quantitative Correlations of Substituent Effects (Part B) – The Hammett Equation
  30. Hammett, L.P. (1937). „The Effect of Structure upon the Reactions of Organic Compounds. Benzene Derivatives”. J. Amer. Chem. Soc. 59 (1), 96–103. o. DOI:10.1021/ja01280a022.  
  31. 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. o. DOI:10.1021/cr00002a004.  
  32. 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)” 69 (12), 2497–2510. o. DOI:10.1351/pac199769122497.  
  33. Pine, S.H., Hendrickson, J.B.; Cram, D.J.; Hammond, G.S.. Organic chemistry. McGraw–Hill (1980). ISBN 0070501157  Section 6-2: Structural Effects on Acidity and Basicity
  34. 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., 723–724. o, 723. o. DOI:10.1039/C19680000723.  
  35. Alder, R.W. (1989). „Strain Effects on Amine Basicities”. Chem. Rev. 89, 1215–1223. o. DOI:10.1021/cr00095a015.  
  36. a b Goldberg, R., Kishore, N.; Lennen, R. (2002). „Thermodynamic Quantities for the Ionization Reactions of Buffers” (reprinted at NIST). J. Phys. Chem. Ref. Data 31, 231–370. o. DOI:10.1063/1.1416902.  
  37. Martell, A.E., Motekaitis, R.J.. Determination and Use of Stability Constants. Wiley (1992). ISBN 0471188174  Chapter 4: Experimental Procedure for Potentiometric pH Measurement of Metal Complex Equilibria
  38. Leggett, D.J.. Computational Methods for the Determination of Formation Constants. Plenum (1985). ISBN 0306419572 
  39. 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), 699–641. o. DOI:10.1016/S0731-7085(98)00010-7.  
  40. Szakács, Z., Hägele, G. (2004). „Accurate Determination of Low pK Values by 1H NMR Titration”. Talanta 62, 819–825. o. DOI:10.1016/j.talanta.2003.10.007.  
  41. 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-Vavelength UV and NMR pH Titrimetric Study of the Micro-Speciation of SKI-606”. J. Pharm. Biomed. Anal. 47 (2), 303–311. o. DOI:10.1016/j.jpba.2008.01.015.  
  42. 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, 1041–1052. o. DOI:10.1007/s00216-003-2020-0.  
  43. Onufriev, A., Case, D.A; Ullmann G.M. (2001). „A Novel View of pH Titration in Biomolecules”. Biochemistry 40, 3413–3419. o. DOI:10.1021/bi002740q.  
  44. 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. o. DOI:10.1021/bi00866a011.  
  45. Dunn, M.J.. Gel Electrophoresis: Proteins. Bios Scientific Publishers (1993). ISBN 187274821X 
  46. Martin, R.. Gel Electrophoresis: Nucleic Acids. Bios Scientific Publishers (1996). ISBN 1872748287 
  47. Brenner, B.M. (Editor), Stein, J.H. (Editor). Acid–Base and Potassium Homeostasis. Churchill Livingstone (1979). ISBN 0443080178 
  48. Scorpio, R.. Fundamentals of Acids, Bases, Buffers & Their Application to Biochemical Systems. Kendall/Hunt Pub. Co. (2000). ISBN 0787273740 
  49. Beynon, R.J., Easterby, J.S.. Buffer Solutions: The Basics. Oxford: Oxford University Press (1996). ISBN 0199634424 
  50. Perrin, D.D., Dempsey, B.. Buffers for pH and Metal Ion Control. London: Chapman & Hall (1974). ISBN 0412117002 
  51. Garfin, D.(Editor), Ahuja, S. (Editor). Handbook of Isoelectric Focusing and Proteomics. Elsevier (2005). ISBN 0120887525U 
  52. Hulanicki, A.. Reactions of Acids and Bases in Analytical Chemistry, Masson, M.R. (translation editor), Horwood (1987). ISBN 0853123306 
  53. Eyal, A.M (1997). „Acid Extraction by Acid–Base-Coupled Extractants”. Ion Exchange and Solvent Extraction: A Series of Advances 13, 31–94. o.  
  54. Avdeef, A.. Absorption and Drug Development: Solubility, Permeability, and Charge State. New York: Wiley (2003). ISBN 0471423653 
  55. Beck, M.T., Nagypál, I.. Chemistry of Complex Equilibria. Horwood (1990). ISBN 0853121435 
  56. van Leeuwen, C.J., Hermens, L. M.. Risk Assessment of Chemicals: An Introduction. Springer, 254–255. o. (1995). ISBN 0792337409 
  57. Skoog, D.A, West, D.M.; Holler, J.F.; Crouch, S.R.. Fundamentals of Analytical Chemistry, 8th, Thomson Brooks/Cole (2004). ISBN 0-03-035523-0  Chapter 9-6: Acid Rain and the Buffer Capacity of Lakes
  58. Stumm, W., Morgan, J.J.. Water Chemistry. New York: Wiley (1996). ISBN 0471051969 
  59. Snoeyink, V.L., Jenkins, D.. Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. New York: Wiley (1980). ISBN 0471511854 
  60. Millero, F.J.. Chemical Oceanography, 3rd, London: Taylor and Francis (2006). ISBN 0849322804 
  61. Millero, F.J., Liu, X. (2002). „The Solubility of Iron in Seawater”. Marine chemistry 77, 43–54. o. DOI:10.1016/S0304-4203(01)00074-3.  
  62. Speight, J.G.. Lange's Handbook of Chemistry, 18th, McGraw–Hill (2005). ISBN 0071432205  Chapter 8

Further reading

  • Albert, A., Serjeant, E.P.. The Determination of Ionization Constants: A Laboratory Manual. Chapman & Hall (1971). ISBN 0412103001  (Previous edition published as Ionization constants of acids and bases. London (UK): Methuen (1962) )
  • Atkins, P.W., Jones, L.. Chemical Principles: The Quest for Insight, 4th, W.H. Freeman (2008). ISBN 1-4292-0965-8 
  • Housecroft, C.E., Sharpe, A.G.. Inorganic chemistry, 3rd, Prentice Hall (2008). ISBN 0131755536  (Non-aqueous solvents)
  • Hulanicki, A.. Reactions of Acids and Bases in Analytical Chemistry. Horwood (1987). ISBN 0853123306  (translation editor: Mary R. Masson)
  • Perrin, D.D., Dempsey, B.; Serjeant, E.P.. pKa Prediction for Organic Acids and Bases. Chapman & Hall (1981). ISBN 041222190x 
  • Reichardt, C.. Solvents and Solvent Effects in Organic Chemistry, 3rd, Wiley-VCH (2003). 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.. Fundamentals of Analytical Chemistry, 8th, Thomson Brooks/Cole (2004). ISBN 0-03-035523-0 

Külső hivatkozások

Extensive bibliography of pKa values in DMSO, acetonitrile, THF, heptane,1,2-dichloroethane, and in the gas phase.
All-in-one freeware for pH and acid-base equilibrium calculations and for simulation and analysis of potentiometric titrationcurves with spreadsheets.
Includes a database with aqueous, non-aqueous, and gaseous phase pKa values than can be searched usingSMILES or CAS registry numbers.
pKa values for various acid and bases. Includes a table of some solubility products.
Explanations of the relevance of these properties to pharmacology.