You are on page 1of 52

PREVIEW

Electrochemistry is concerned with the properties of due mainly to Debye and HUckel. This theory focuses
solutions of electrolytes and with processes that occur at attention on the distribution of positive and negative
electrodes. A useful starting point is provided by Fara- ions in solution as a result of the electrostatic forces. In
day's laws of electrolysis, which relate the mass of a the near neighborhood of each ion there are more ions
substance deposited at an electrode to the quantity of of opposite sign than of the same sign, and it is
electricity (current X time) passed through the solution, convenient to speak of an ionic atmosphere as existing
and to the relative atomic or molecular mass of around each ion. The thickness of the ionic atmosphere
the substance. decreases as the concentration is raised. When an ion
Important properties of solutions of electrolytes are moves through the solution as a result of an applied
the resistance, the conductance, and the electrolytic potential field, the atmosphere lags behind and causes a
conductivity. To compare solutions of different retardation of the motion; this is known as the
concentrations, it is convenient to introduce the molar relaxation or the asymmetry effect. The moving ionic
conductivity A (formerly called the equivalent atmosphere drags solvent molecules with it, and as a
conductivity), which is the conductivity when one mole result there is an additional retardation; this is known
of the electrolyte is in between two electrodes that are a as the electrophoreric effect.
standard distance apart. Kohlrausch's law of independent migration of ions
Different theories are required for solutions of states that the molar conductivity of an electrolyte is the
weak electrolytes, which are only slightly dissociated, sum of the individual ionic conductivities. The mobility
and for strong electrolytes, which are completely of an ion is the speed with which it moves in a unit
dissociated. The former are satisfactorily treated by a potential gradient, and it is proportional to the ionic
theory due to Arrhenius. which considers how the conductivity. The transport number of an ion is the
dissociation equilibrium shifts as the concentration of fraction of the current carried by that ion. Transport
the electrolyte is varied. The theory finds its numbers can be measured experimentally, and they
quantitative expression in Osrwald's dilution law, allow the molar conductivity to be split into the
which relates the molar conductivity A at any individual ionic conductivities.
concentration to that at infinite dilution A. The ratio This chapter is also concerned with the
AfA is the degree of dissociation a. thermodynamics of ions in solution, and several
This theory does not apply to strong electrolytes, important matters are involved: the enthalpies and
which are dealt with by an entirely different treatment entropies of hydration of ions and the activity

263
coefficients of ions. The latter are treated in an - Another important matter, particularly in biological
approximate manner by the Debye-Huckel limiting law, systems, is the distribution of electrolytes across
according to which the logarithm of the activity membranes that are permeable to some ions but not to
coefficient y, of an ion of the ith type is proportional to others. This problem is treated by the theory of the
the square root of the ionic strength of the solution. Donnan equilibrium.

OBJECTIVES

After studying this chapter, the student should be able to: Explain the properties of ions in solution, including
the special properties of the hydrogen ion in aqueous
Understand Faraday's laws of electrolysis.
solution.
Define molar conductivity.
Understand how the Debye-Htickel theory interprets
Explain the Arrhenius theory of strong electrolytes the activity coefficients of ions in solution.
and derive Ostwald's dilution law.
Explain the concept of the ionic product.
Describe the ionic atmosphere and how it affects the
Understand solubility products and the theory of the
conductivity of a solution of an electrolyte (the
Dorman equilibrium.
Debye-Mickel theory).
Understand the concept of a transport number and
how it is measured.

264
At the atomic level, electrical factors determine the structures and properties of
chemical substances. The important branch of physical chemistry known as
electrochemistry is particularly concerned with the properties of solutions of
electrolytes and with processes occurring when electrodes are immersed in these
solutions. This chapter deals with the electrochemistry of solutions, a topic that is
of special interest since studies in this area, in the late nineteenth century, played an
important role in the development of physical chemistry.
Much can be learned about the behavior of ions in water and other solvents by
investigations of electrical effects. Thus, measurements of the conductivities of
aqueous solutions at various concentrations have led to an understanding of the ex-
tent to which substances are ionized in water, of the association of ions with
surrounding water molecules, and of the way in which ions move in water.
It is convenient to begin our study of electrochemistry with a brief survey of
the development of concepts related to electricity. It has been known since ancient
times that amber, when rubbed, acquires the property of attracting light objects
such as small pieces of paper or pitch. In 1600 Sir William Gilbert (1544-1603)
coined the word electric, from the Greek word for amber (elektron), to describe
substances that acquired this power to attract. It was subsequently found that mate-
rials such as glass, when rubbed with silk, exerted forces opposed to those from
amber. Two types of electricity were thus distinguished resinous (from substances
like amber) and vitreous (from substances like glass).
In 1747 the great American scientist and statesman Benjamin Franklin
(1706-1790) proposed a "one-fluid" theory of electricity. According to him, if bod-
ies such as amber and fur are rubbed together, one of them acquires a surplus of
"electric fluid" and the other acquires a deficit of "electric fluid." Quite arbitrarily,
Franklin established the convention, still used today, that the type of electricity pro-
duced on glass rubbed with silk is positive; conversely, the resinous type of
electricity is negative. The positive electricity was supposed to involve an excess of
electric fluid: the negative electricity, a deficit. We now know that negative electric-
ity corresponds to a surplus of electrons, whereas positive electricity involves a
deficit. Franklin's convention, however, is now so well entrenched that it would be
futile to try to change it.
Until 1791 the study of electricity was entirely concerned with static electric-
ity, in the form of charges developed by friction. In that year Luigi Galvani
(1737-1798) brought the nerve of a frog's leg into contact with an electrostatic ma-
chine and observed a sharp convulsion of the leg muscles. Later he found that the
twitching could be produced simply by bringing the nerve ending and the end of
the leg into contact by means of a metal strip. Galvani believed that this effect
could only be produced by living tissues. Then in 1794 the Italian physicist
Alessandro Volta (1745-1827) discovered that the same type of electricity could be
Voltaic Pile produced from inanimate objects. In 1800 he constructed his famous "Voltaic pile,"
which consisted of a number of consecutive plates of silver and zinc, separated by
cloth soaked in salt solution. From the terminals of the pile he produced the shocks
and sparks that had previously been observed only with frictional machines. With
the experiments of Galvani and more particularly of Volta, the subject of electricity
entered into a new era in which electric currents played predominant roles.
Very shortly after Volta carried out his experiments, the British scientific writer
William Nicholson (1753-1815), in collaboration with the surgeon Anthony Carlisle
(1768-1842), decomposed water by means of an electric current and observed that
Electrolysis oxygen appeared at one pole and hydrogen at the other. The word electrolysis (Greek

265
266 Chapter 7 Solutions of Electrolytes

Iv.vi,v, seth rig free) was later used by' the Fnglisli scientist Michael Faraday
C 1791-1807) to describe such splitting by an electric current. Faraday gave the nanie
In an electrolytic cell, the negative cathode (Greek eathados, way down) to the electrode at which the hydrogen collects
electrode is called the cathode, and anode (Greek aita, up: hodos, way) to (he other electrode. The important studies
and the positive electrode the of Faraday, which put the subject ol electrolysis on a quantitative basis, arc consid-
anode; these terms were ered laici.
introduced by Faraday in 1834. In
a cell that generates electricity
(Chapter 8), the opposite is true;
the cathode is the positive Electrical Units
electrode and the anode the
negative electrode. To avoid The clectroMoNc 'e I' between two chari,es (j and Q , separated by it distance r

confusion it is best to specify the in it vaeuuni i


charges, plus or minus.

The constant E, is the pernhiIlil'i!v a/a t'a('Uwi and has the value 8.1154 X 10 1 2
I 1 in '. The factor 47r is introduced into this expression in order that the Gauss
and Poisson L'qualit)Ils (see pp. 278-279 ) are free of' this lictoi'. if' the char-'es are
Dielectric Constant in it nieditum having it re/u!iI'e /wrmiI!i1'uv. or dielectric con.slaill. or . the eqUation
for the force is

12
F = (7.2)
4ir,r

For example. water at 2 '(' has a dielectric constant of about 7. with the result
that the electiosiut ic forces between ions are reduced by this factor. The SI twit of
charge Q is the coulouth. C. and that of' distance i' is the metre. in. The unit of force
F is the newton. N. which is joule per metre. J in
The eh'i 'ii'h lie/il F at any point is the force exerted on it unit charge (I C') at
that point. The held strength at it distance r from a charge Q. in it medium of dielec-
tric coilsiant . is thus

Q .' (7.3)
4 1 r

I
The SI unit of field strength is therefore newton per coulomb (N C
= J C in
However, siiicc joule = volt coulomb = newton metre (see Appendix A), the SI
(hut of' hicki sti'englh is usually expressed as volt per metre (V ii
Just as it mechanical force is the negative gradient of it potential. the ('le('frie
field sire,,giI, is the negative gradient of an electric potential 1 . Thus (lie field
strength F at a distance r t'rom a charge Q is given by

(7.4)

and the electric potential is thus

'Sc' A11'nthi A For it further discussion of the ICrni 47T.


7,1 Faraday's laws of Electrolysis 267

Q
(7.5)
4E0fr

The SI unit of the electric potential is the volt (V).2


Further details about electrical Units, including a comparison of SI units with
the electrostatic units, are to be found in Appendix A.

7.1 V, Faraday's Laws of Electrolysis


During the years 1833 and 1834 Michael Faraday published the results of an ex-
tended series of investigations on the relationships between the quantity of electricity
passing through a solution and the amount of material liberated at the electrodes. He
formulated Faraday's laws of electrolysis, which may be summarized as follows:
1. The mass of an element produced at an electrode is proportional to the
quantity of electricity Q passed through the liquid; the SI unit of Q Is the
coulomb (C). The quantity of electricity Is defined as equal to the current
I (SI unit = ampere, A) multiplied by the time I (SI unit = second, s):
Q=It (7.6)
2. The mass of an element liberated at an electrode Is proportional to the
equivalent weight of the element.
The SI recommendation, supported by the International Union of Pure and Ap-
plied Chemistry and other international scientific bodies, is to abandon the use of
the word equivalent and to refer to moles instead. However, the concept of the
equivalent is still employed, even though the name has been dropped. Suppose, for
example, that we pass 1 A of electricity for 1 h through a dilute sulfuric acid solu-
tion and through solutions of silver nitrate, AgNO3, and cupric sulfate, CuSO4.3
The masses liberated at the respective cathodes are
0.038 g of H2 (relative atomic mass, A,. = 1.008;
relative molecular mass, M,. = 2.016)
4.025 g ofAg (A r = 107.9)
1.186 g of Cu (A,. = 63.6)
These masses 0.038,4.025, and 1.186 are approximately in the ratio 1.008/107.9/31.8.
Therefore, the amount of electricity that liberates 1 mol of Ag liberates 0.5 mol of H2
and 0.5 mol of Cu. These quantities were formerly referred to as 1 equivalent, but the
modern practice is to speak instead of 1 mol of +H2 and 1 mol of --Cu. These are the
quantities liberated by 1 mol of electrons:
e+H - 4H2 and e + Cu2+ - -Cu
In what follows, when we refer to I mol we may mean I mol of a fraction of an entity
(e.g., to -Cu or -SO).

21n terms of base SI units, volt = kg m2 s 3 A 1


3The symbol for hour is h.
268 Chapter 7 Solutions of Electrolytes

MICHAEL FARADAY'(17 cU367) 5


chemistry rather than physics. In 1831 he discovered
the phenomenon of electromagnetic induction, and in
the following year he began to investigate whether the
electricities produced in various ways were identical. It
Is interesting to note that he did not obtain his two laws
of electrolysis empirically; he deduced them from what
he discovered about the nature of electricity, and
proceeded to verify them.
Faraday was ignorant of mathematics beyond
simple arithmetic, but he developed qualitative theories.
In Faraday's view an electric current is due to a buildup
and breakdown of strain within the particles of matter
through which the current flows. His ideas about
electricity and magnetism led to the establishment of
Faraday was born in Newington, Surrey, which is not far modern field theory. In the 1850s and 1860s Clerk
from London. His father was a blacksmith and the Maxwell (1831-1879) formulated Faraday's Ideas into a
family was of very limited means. At the age of 14 mathematical theory of electromagnetic radiation.
Faraday was apprenticed to a bookbinder, and he took Faraday was deeply religious, being a loyal member
the advantage of reading many of the books he bound. of the Sandemanian Church. He modestly declined all
His interest in science was particularly aroused by Jane honors and avoided scientific controversy. He preferred
Marcet's famous Conversations on Chemistry, which to work alone and without an assistant, and as a result
first appeared in 1809. he founded no school of research. He became famous
Faraday was fortunate to obtain the position of for his Friday Evening Discourses, which he founded at
laboratory assistant to Sir Humphry Davy (1778-1829), the Royal Institution and of which he gave over a
the Director of the Royal Institution. In 1825 he hundred himself. One of them was his well-known
succeeded Davy as director of the Royal Institution Christmas lecture on the Chemlcal History of a Candle."
laboratories, and in 1833 he also became Fullerian
Professor of Chemistry, a post that was created for him
References: L. Pierce Williams, Dictionary of Scientific
by John Fuller, a Member of Parliament. Fuller had Biography, 4, 527-540(1971).
attended many of the public lectures at the Royal D. K. C. Macdonald, Faraday, Maxwell and Kelvin,
Institution, and had often slept through them; his New York: Doubleday & Co., 1964.
endowment was in gratitude for the peaceful hours Sir George Porter, 'Michael Faraday--Chemist" (Faraday
Lecture of the Chemical Society), Proc, Roy. Institution, 53,
thus snatched from an otherwise restless life."
90-99(1981).
After 1821 much of Faraday's work was on electricity L. Pierce Williams, Michael Faraday, A Biography, New
and magnetism, which at the time was considered to be York: Basic Books, 1965.

The proportionality factor that relates the amount of substance deposited to the
quantity of electricity passed through the solution is known as the Faraday con-
stant and given the symbol F. In modern terms, the charge carried by I mol of ions
bearing z unit charges is zF. According to the latest measurements, the Faraday
constant F is equal to 96 485.309 coulombs per mole (C mot - '); in this discussion
the figure will be rounded to 96 485 C mol '. In other words, 96 485 C will liberate
I mot of Ag, I mot of +H2 (i.e., 0.5 mot of H 2 ), etc. If a constant current I is passed
for a period of time 1, the amount of substance deposited is It/(96 485 C mol).
It follows that the (negative) charge on I mot of electrons is 96 485 C. The
charge on one electron is this quantity divided by the Avogadro constant L and is thus

96 485 C moF'
= 1.602 x 10- ' C
6.022 X 1023 moF
7.2 Molar Conductivity 269

EXAMPLE 7.1 An aqueous solution of gold (III) nitrate, Au(NO3) 3 , was elec-
trolyzed with a current of 0.0250 A until 1.200 g of Au (atomic weight 197.0) had
been deposited at the cathode. Calculate (a) the quantity of electricity passed, (b)
the duration of the experiment, and (c) the volume of 0, (at NSTP 4 ) liberated at
the anode. The reaction at the cathode is
3e + Au flt - Au or e + --Au +
The equation for the liberation of 02 is
2H20 -402 + 4W + 4e or +H20 4 +H+e
Solution a. From the cathode reaction. 96 485 C of electricity liberates I mol
of 4Au (i.e., 197.0/3 g). The quantity of electricity required to produce 1.200 g of
Au is thus
96485 x 1.200
l.76X 103 C
197.013
b. Since the current was 0.0250 A, the time required was
1.76X 103(As) 104
7.4X
0.025(A)
c. From the anode reaction. 96 485 C liberates Imol 02, so that 1.76 X 103 C
produces
1.76 X 103
= 4.56 X 10 mol O
96485X4
The volume of this at NJSTP is
4.56 X 10 (mol) X 24.8(dm moL') = 0.113 dm3

Faraday's work on electrolysis was of great importance in that it was the first
to suggest a relationship between matter and electricity. His work indicated that
atoms, recently postulated by Dalton, might contain electrically charged particles.
Later work led to the conclusion thai an electric current is a stream of electrons and
that electrons are universal components of atoms

7.2 Molar Conductivity


Important information about the nature of solutions has been provided by measure-
ments of their conductivities. A solution of sucrose in water does not dissociate into
ions and has the same electrical conductivity as water itself. Such substances are
known as ,ioneleetmlvte.s. A solution of sodium chloride or acetic acid, on the other
hand, forms ions in solution and has a much higher conductivity. These substances
are known as elect wi vtes.

New standard tenipennure and pres,lre. vi,.. 25.0 C and I bar: at NSTP I rnol of gas occupies 24.8 din.
270 Chapter 7 Solutions of Electrolytes

Sonic electrolytes in aqueous sol Ut ions at the same niolal i ty conduct better
than others. This suggests that some electrolytes es isi as ions in solution to it
rea1cr extent than do others. More detailed investications have indicated that cer-
taiii substances such as the salts sodium chloride and cupric sulfate, as well as some
acids such as hydrochloric acid. OCCUr almost entirely as ions when they are ill
Strong and Weak aqueous solution. Such substances are known as strong electrolytes. Other sub
Electrolytes stances, including acetic acid and ammonia. are present only partially as ions
(e.g., acetic acid exists ill solution partly as CHCOOH and partly as CH 1C0() 4-

H ). These substances are known as weak electrolytes.


Further insight into these matters is provided by studies al the way in which
the electrical canduclivities of solutions v;iiv with the concentration of solute.
Accordine to Ohm's law, the resistance R of a slab of material is equal to the
electric potential difference V divided by the electric current I:

1? =- - ( 7.7)

The SI unit of potential is the volt (V) and that lw current is the ampere (A). The
unit of electrical resistance is the ohm." given the symbol f (omega). The recipro-
cal of the resistance is the e/ec'Iru(il (onth:Uun(e CL the SI unit of which is the
ie,wm (5 ) The electrical conductance of' material of length I and cross-
sectional area 4 is proportional to A and inversely proportional to I:

G (conductance) = K (7.)

The proportionality constant K. which is the conductance 01- a unit cube (see Figure
7.1 a). is known as the eomfueiiritv C lt)rIlieI-l\ as the specific conductivity): ln' the par-
Electrolytic Conductivity ikuharcase ala solution of an elecirolyte. it is known as the elcctrolylk COfldUCtiVit.
Its SI unit is D I Ill ( Sm hut the unit more commonly used is 11 1 cm
I)
(S cm
The electrolytic conLluctivity is not it suitable quant ii>' for conipari ng the con-
duclivilies at dilThreni solutions. II' a solution of one electrolyte is much iiioi-e
concentrated than another, it may have a higher conductivity simply because it con-
tuifls more ions. What is needed instead is it property in which there has been some
conipensat ion for the dilieience in concentrations. In the 1880s the German physi-
cist Friedrich Wilhelm Georg Kohlrausch (I 840-19 10) clarified the early theory by
Molar Conductivity employing the concept of what is now called the molar conductiityt and given
the symbol A I lanihda. Ii is defined as the electrolytic conductivity K divided by
concentration e.

A= (7.9)

The meaning of A can he visualized as follows. Suppose that we construct a cell


having parallel plates a unit distance apart. the plates being at' such an area that tar

In ternis of tsic SI IIIuII. 1)11111 ki 111 A


Ii Wds lorinerly Lulled (tiC CLILIIVJICI(t CL)I1(ILIC(Ic.iIy. IiLII ILIS tli;idvnoted) IURAC no lUIL!Cr rLCl,Ii1I1IeIILl
the Use Of 111C word cq11Iv,IILI1l. which is sonminies iiiihiu,iii': An cntiiclv LIlilerelli equlvulLill is
invoked it a subsiance jCis as all IESidLtiili( LW teLIUciilL, LILCIII. lilsiCud ot -.pcakine iii an eqtii\ukiil of
C LI S( 14 . 11W example. which reqLuIIL's us II) know that we mean hail a tuttle ol C ti S( ) . we speak of a inole

iLl .CiiSO.
7.2 Molar Conductivity 271

Conductance, G = Molar conductivity. =

Volume= imI

Unit separation

Area =A Unit cube


/'Unit cube

FIGURE 7.1
(a) The relationship between con- Length = I
ductance and electrolytic conduc- Unit area Electrolytic conductivity, ,c
tivity. (b) The relationship between Electrolytic
molar conductivity and electrolytic conductivity, h

conductivity. a. b.

iir(jcuit' solution I niol ut elceirulvle (e... HCI. CuSO 1 ) is present in it (see


Fiurc 7.1 h). 'l'lic molar conductivity A is the conductance (i.e., the i'eciprocal of
the resistance) across the plates. We iieed not actually construct cells For each solu-
tion lr which we need to know the molar conduct i' i. since we can calculate it
ftoiu the electrolytic conductivity K. II' the eunceiutratioml of Ihe solution is e. the
Volume of, the hypothetical molar conductivity cell iiiiist he (I imiul i/(' in order for
I ninl 10 he between the pIat 4 Iure 7. I hI. The 11101,tr conductivity A is themelote
the electrohtic conducliv liv K multiplied hy I Ir. In using l.ti. 7.9. care must he
taken ith thus, as shown hy l:xasuj)le 7.2.

IXAiIPLL' 7.2 The electrolytic conductivity of' a (3.1 M solution of acetic acid
(c)rrected for the conductivity of the alct') was bLind to he 5.3 X 10
ciii . Calculate the molar conductivity.

Solution From Ftj. 7.9.

5.3 If t ml - 0 'cni )
0 =
Potassium chloride I). I (nuol dnu ) It) 'I nuol cm
- (strong electrolyte) I
5 Q rnul
100
>.. This result can also he written as 53 S ciii mmd 1 . where S stands for siemens,
> the unit of, conductance, which is equivalent In the reciprocal ohm,
U

o
c 50
Acetic acid The inlporiiilice ol the imiolat colikiticlix Ity is that it tivcs inlmmnatitimi about the
weak electrolyte)
conductis ily ul the ions produced in solutioti by I mitol of a substance. In all cases
the niokir cunducii ily diminishes as hue concentration is raised, and Iwo patterns
Concentration/mot dm : ui heh;ivmor can be distmiwu i shed. From Fitttre 7.2 we see that tr st oiii dcc-
im'olytes dic ittolar conductivity halls only sliththv as the concentration is raised. On
FIGURE 7.2
the other hand. the weak elccti'olvles produce lever lons. and exhibit a nitich more
The variations of molar conductiv-
ity A with concentration for strong Promuottiucell I'.-ill oh' A with incicasi ni concentration. as shown by the lower curve in
and weak electrolytes. Figure 7.2.
272 Chapter 7 Solutions of Electrolytes

We can extrapolate the curves hack to 1cm concentration and obtain a qLlaultity
kno' ii as A', the niolur (on(Irt(tO'itv at infimt(' thIuJiw/, or zero concentration. With
weak electrolytes this extrapolation is unreliable, and an indirect method, explained
oil pp. 291-292, is usually employed for its calculation.

7.3 Weak Electrolytes: The Arrhenius Theory


The conductivity (Lila of' Koli lrauscli. and the tact that the beat of neutralization
01 a stioiW acid by a Strong base in dilute aqueous solution was pi't ically the
same Ioi all s(rolliZ acids and Stroll-' (about 54.7 kJ niol 1 at 25 ':C), led
the Swedish scientist Svanlc Auuust Arrhcnius (1859-1927) to propose in 18 871 a
new theory nb electrolyte solutions. According to this theory, there exists an
equilibnuin in solution between undissociated molecules AR and the ions A
and B

AR A' -I- B (7.10)

Al very low concentrations this equ i Ii briuni lies over to the right, and the molar
conductivity is close to A''. As the concentration is increased, this equilibrium shills
to the Icit and the molar conductivity decreases troni A to a howet' value A. The
Degree of Dissociation degree of dissociation, that is. the fraction of AR in the lrni A + B - is A/A",
which is denoted by the symbol 1r:

A
7.11)
A"

At high concentrations there is less dissociation, while at infinite dilution there is


complete dissociation and the molar conductivity then is A.
'['his theory explains the constant heat ol neutralization of all strong acids and
bases. which are dissociated In a considerable exteni . the neutralizatiomi involving
simply i lie react ion H -F- Oil > HO. However, as we shall see (Section 7.4).
Arrhenius's explaiiatioii of, the decrease in A with increasing concentration is valid
only for weak electrolytes: another explanation is required lor strong electrolytes.
Soon after Arrheni us's theory was proposed. van'! Hoff Found that osmotic
pressure incas tire men Is that lie had made gave it considerable support. Va mit Hull'
had lou id that the osmotic pressures ol solutions ot electrolytes were always
considerably higher than predicted by the osmotic pressLti'c equation for nonelec-
i-olytes (Eq. 5.134). He proposed the modified cci nat ion

icRT (7.12)

Van't Hoff Factor where i is known as the 'an't Hoff factor. For sti'orit electrolytes, the vant Hoff
factor is approximately equal to the number 01' ions formed from one molecule: thus
for NaCl, HCI, etc., I = 2: ['or Na,SO 4, RaCE, etc., I = 3: and so forth. However,
the value of i decreases with increasing ionic strength. For weak electrolytes the
van'! Hoff factor I involves the degree of dissociation. Suppose that one fliOlecLi]e of
a weak electrolyte would produce. if there were complete dissociation, v ions. The
number Of ions actually produced is thus r', and the number of undissociated mole-
cules is I - , so that the total number of' particles produced from I molecu]e is

a + pa (7.13)
7.3 Weak Electrolytes: The Arrhenius Theory 273

SVANTE AUGUST ARRHENIUS (1859-1927)


director of physical chemistry at the Nobel Institute in
Stockholm.
Arrhenius continued for many years to work on
electrolytic solutions, but to the end of his life never
accepted the fact that strong electrolytes are
completely dissociated, and that their behavior is
strongly influenced by interionic forces (Section 7.4).
He worked in a variety of other fields, and made
important contributions to immunochemistry,
cosmology, the origin of life, and the causes of the ice
age. It was he who first discussed what is now called
the 'greenhouse effect." He published many papers
and a number of books, including Lehrbuch der
kosmischen Physik (1903), Immunochemistry (1907),
Arrhenius was born in Vik (Wijk), which is near and Quantitative Laws in Biological Chemistry (1915).
Uppsala, Sweden. He was a student at the University Arrhenius received the 1903 Nobel Prize in
of Uppsala but carried out his doctoral research in chemistry, the citation making reference to the "special
Stockholm under the physicist Erik Edlund. His thesis, value of his theory of electrolytic dissociation in the
completed in 1884, was on the conductivities of development of chemistry."
electrolytic solutions. It did not suggest the dissociation
of electrolytes into ions, and received a poor rating
References: H. A. M. Snelders, Dictionary of Scientific
from the examiners. Arrhenius then had discussions
Biography, 1, 296-302(1970).
with Ostwald, van't Hoff, and others, and as a result J. W. Servos, Physical Chemistry from Ostwald to
postulated his theory of electrolytic dissociation in Pauling: The Making of a Science in America, Princeton
1887. University Press, 1990.
After holding some teaching positions, in 1895 K. J. Laidler, The World of Physical Chemistry, New York:
Oxford University Press, 1993.
Arrhenius was appointed professor of physics at the
Elisabeth Crawford, Arrhenius: From Ionic Theory to the
University of Stockholm; he served as its Rector from Greenhouse Effect, Science History Publications, Canton,
1897 to 1902. From 1905 until his death he was Ohio, 1998,

Therefore

(7.14)

It is thus possible to calculate a from osmotic pressure data. For weak electrolytes
the values so obtained are in good agreement with A/A. and the Arrhenius theory
therefore applies satisfactorily to such systems.

Ostwald's Dilution Law


Wilhelm Ostwald received the
1909 Nobel Prize in chemistry In 1888 the ideas of Arrhenius were expressed quantitatively by Friedrich Wilhelm
for his wide-ranging research on Ostwald (1853I 32) in terms ofa (h/anon law. Consider an electrolyte AB that exists
catalysis, chemical equilibrium, in solution partly as the undissuciated species AB and partly as the ions A and B
and rates of chemical reactions.
AB + B
He was very concerned with
energy waste, writing "Dissipate The equilibriiini constant, on the assumption of ideal behavior, is
no energy, but strive to use
energy by converting it into more K, =
113 I (7.15)
useful forms." JABI
274 Chapter 7 Solutions of Electrolytes

Suppose that an amount n of the electrolyte is present in a volume V and that the frac-
tion dissociated is a; the fraction not dissociated is 1 - a. The amounts of the three
species present at equilibrium, and the corresponding concentrations, are therefore
AB
Amounts present at equilibrium: n(l - a) na na
n(l a) na na
Concentrations at equilibrium:

The equilibrium constant is


(na/V)2 na2
K.= = (7.16)
[n(1 a)]IV V(l a)
Therefore, for a given amount of substance the degree of dissociation a must vary
with the volume V as follows:
a
= constant X V (7.17)
1a
Alternatively, since n/V = c, we can write
Ca
=K (7.18)
I a
The larger the V, the lower the concentration c and the larger the degree of dissocia-
tion. Thus, if we start with a solution containing 1 mol of electrolyte and dilute it,
the degree of dissociation increases and the amounts of the ionized species in-
crease. As V becomes very large (the concentration approaching zero), the degree
of dissociation a approaches unity; that is, dissociation approaches 100% as infinite
dilution is approached. The experimental value of A corresponds to complete dis-
sociation; at finite concentrations the molar conductivity A is therefore lower by the
factor a(= A/A). The dilution law can thus be expressed as

KV= (7.19)
1(A/A)
or

= c(A/A0)2
K (7.20)
1 - (A/A)

Equation 7.20 has been found to give a satisfactory interpretation of the variation of
A with c for a number of weak electrolytes, but for strong electrolytes there are se-
rious deviations, as discussed in the next section.

Strong Electrolytes
Arrhenius maintained all his life that his theory applied without any modification to
strong as well as to weak electrolytes. However, some of the results for strong elec-
trolytes were soon found to be seriously inconsistent with the theory and with
Ostwald's dilution law. For example, the "constants" K calculated from A/A val-
7,4 Strong Electrolytes 275

ues on the basis of Ostwald's dilution law (Eq. 7.20), were found for stronger acids
and bases to vary, sometimes by several powers of 10, as the concentration was var-
ied. Furthermore, values for the conductivity ratio A/A were sometimes
significantly different from those obtained from the van't Hoff factor (Eq. 7.12).
We saw earlier that the constancy of heats of neutralization of strong acids and
bases was originally regarded as good evidence for the Arrhenius theory. But the
heats were too constant to be consistent with the theory, because at a given concen-
tration there should have been small but significant differences in the degrees of
ionization of acids such as HCI, HNO, and H,SO4, and these differences should
have given measurable differences between the heats of neutralization. These dif-
ferences could not be observed, a result that suggests that these acids are
completely dissociated at all concentrations.
Strong support for this idea was obtained between 1909 and 1916 by the Dan-
ish physical chemist Niels Janniksen Bjerrum (1879-1958), He examined the
absorption spectra of solutions of a number of strong electrolytes and found no evi-
dence for the existence of undissociated molecules. The fall in A with increasing
concentration must therefore, for strong electrolytes, be attributed to some cause
other than a decrease in the degree of dissociation.
cl- These anomalies led a number of scientists to reject Arrhenius's theory com-
t pletely, and to conclude that ions are not present free in solution. One of the most
Cl- _;. Cl- vigorous opponents of electrolytic dissociation was the British scientist Henry Ed-
/ (" / ward Armstrong (1848-1937), who even went so far as to lampoon the idea in two
Cl- - cr fairy tales, entitled A Dream of Fair Hydrone and The Thirst of Salt Water His main
objection to the theory was that it appeared to give the solvent merely a passive
Cl- rolein his expressive phrase, water was "merely a dance floor for ions." He pre-
ferred a solvation theory, in which the water played a more active role, There was
a.
indeed some substance to his criticism, but Armstrong and others were wrong to try
to overthrow completely the idea that substances can be dissociated into ions in so-
lution. What was needed instead was modification of the idea. Later treatments
N allowed the solvent to play a more active role, and it was realized that solvent mol-
ecules are sometimes bound quite strongly to ions, as will be considered in Section
7.9. The importance of the electrostatic forces between ions was also taken into ac-
/ / count, as will now he discussed.

Debye-Hckel Theory
---- In 1923, the Dutch physical chemist Peter Debye (1884-1966) and the German
physicist Erich l-lckel (1896-1980) published a very important mathematical treat-
ment of the problem of the conductivities of strong electrolytes. The decrease in the
molar conductivity of a strong electrolyte was attributed to the mutual interference of
b. the ions, which becomes more pronounced as the concentration increases. Because
of the strong attractive forces between ions of opposite signs, the arrangement of ions
FIGURE 7.3
The distribution of chloride ions in solution is not completely random. In the immediate neighborhood of any positive
around a sodium ion (a) in the ion, there tend to be more negative than positive ions, whereas for a negative ion
crystal lattice and (b) in a solution there are more positive than negative ions. This is shown schematically in Figure 7.3
of sodium chloride. In the solution for a sodium chloride solution. In solid sodium chloride there is a regular array of
the interionic distances are
greater, and the distribution s not i .
sodium and chloride ions. As seen in Figure 7.3a, each sodium ion has six chloride
i
ions as its nearest neighbors, and each chloride on has six sodium ions. When the
so regular; but near to the sodium
ion there are more chloride ions sodium chloride is dissolved in water, this ordering is still preserved to a very slight
than sodium ions, extent (Figure 7.3b). The ions are much farther apart than in the solid; the electrical
276 Chapter 7 Solutions of Electrolytes

attractions are therefore much smaller and the thermal motions cause irregularity.
The small amount of ordering that does exist, however, is sufficient to exert an im-
portant effect on the conductivity of the solution.
The way in which this ionic distribution affects the conductivity is in brief as
follows; the matter is discussed in more detail later (see Figure 7.6). If an electric
potential is applied, a positive ion will move toward the negative electrode and
must drag along with it an entourage of negative ions. The more concentrated the
solution, the closer these negative ions are to the positive ion under consideration,
and the greater is the drag. The ionic "atmosphere" around a moving ion is there-
fore not symmetrical; the charge density behind is greater than that in front, and
this will result in a retardation in the motion of the ion. This influence on the speed
Relaxation Effect of an ion is called the relaxation, or asymmetry, effect.
A second factor that retards the motion of an ion in solution is the tendency of
the applied potential to move the ionic atmosphere itself. This in turn will tend to
drag solvent molecules, because of the attractive forces between ions and solvent
molecules. As a result, the ion at the center of the ionic atmosphere is required to
move upstream, and this is an additional retarding influence. This effect is known
Electrophoretic Effect as the electrophoretic effect.

dV The Ionic Atmosphere


)01

The Debye-Huckel theory is very complicated mathematically and a detailed treat-


ment is outside the scope of this book. We will give a brief account of the theoretical
/ treatment of the ionic atmosphere, followed by a qualitative discussion of the way in
/ which the theory of the ionic atmosphere explains conductivity behavior.
/ Figure 7.4 shows a positive ion, of charge ze, situated at a point A; e is the
unit positive charge (the negative of the electronic charge) and z is the valence of
/ the ion. As a result of thermal motion there will sometimes be an excess of positive
/ ions in the volume element and sometimes an excess of negative ions. On the aver-
ze age, the charge density will be negative because of the electrostatic forces. In other
words, the probability that there is a negative ion in the volume element is greater
FIGURE 7.4 than the probability that there is a positive ion. The negative charge density will be
An ion in solution of charge ze, greater if r is small than if it is large, and we will see later that it is possible to de-
with an element of volume dV fine an effective thickness of the ionic atmosphere.
situated at a distance r from it. Suppose that the average electric potential in the volume element dV is 4), and
let z and z_ be the positive numerical values of the ionic valencies; for example.
for Na, z = 1; for Cl -, z_ = 1. The work required to bring a positive ion of
charge ze from infinity up to this volume element is ze4), a positive quantity. Be-
cause the central positive ion attracts the negative ion, the work required to bring a
negative ion of charge z_e is ze4), a negative quantity.
The time-average numbers of the positive and negative ions present in the vol-
ume element are given by the Boltzmann principle (see Eq. 1.78, p. 29; the
principle is dealt with in detail in Section 1.11 ): 7
dN = Ne_zTdV (7.21)
and
dN_ = N_ e _z_TdV (7.22)

7 We use roman e here for the exponential function to distinguish it from the charge on the electron e.
7.4 Strong Electrolytes 277

where N, and N_ are the total numbers of positive and negative ions, respectively,
per unit volume of solution; k13 is the Boltzmann constant; and T is the absolute
temperature. The charge density in the volume element (i.e:, the net charge per unit
volume) is given by

= e(z+ dN+ - z_dN_) = e(N 4,/kT - N_z_e2(tT)


+z e (7.23)
dV
For a univalent electrolyte (one in which both ions are univalent) z + and z_ are
unity, and N+ and N_ must be equal because the entire solution is neutral; Eq. 7.23
then becomes

p = Ne(eT - e/T) (7.24)


where N = N, = N_. If e4,/k8T is sufficiently small (which requires that the vol-
ume element is not too close to the central ion, so that 4, is small), the exponential
terms are given by8

e_fT I-- and + (7.25)


kBT kT
Equation 7.24 thus reduces to

Z Electric field =-- (7.26)


strength kT
E= i--
47tE0r2 In the more general case, in which there are a number of different types of ions, the
Electrical expression is
-' potential
e24,
= p ----- N1 z ( 7.27)
.BT i
F6

Charge=Q x
a. where N1 and zi represent the number per unit volume and the positive value of the
valence of the ions of the ith type. The summation is taken over all types of ions
present.
Equation 7.27 relates the charge density to the average potential 4,. In order to
o L7NNomaI to
surface obtain these quantities separately it is necessary to have another relationship be-
E'l/
tween p and 4,.
E cos 9
This is provided in the following way.9 From Eq. 7.3 we have the magnitude E
ds of the electric field at a point P, a distance r from the origin. The direction of the
1i
Solid angle vector field due to a positive charge at the origin is the same as the direction of the
do = dS cos O vector from the origin to the point P (see Figure 7.5a). To make the connection we
will be interested in determining the flux of the electric field (i.e., the number of
fdW 47r lines of force) through an element of area on a surface perpendicular to the electric
field vector. Consider a closed surface S surrounding a charge Q (Figure 7.5b). At
b.

FIGURE 7.5
(a) The electric field strength E
"The expansion of e is
and the electric potential , at a
distance r from a charge Q. e" I
(b)A diagram to illustrate Gauss's and when x is sufficiently small, we can neglect terms beyond x; thus ex I 1- x and e I - x. The
theorem. A charge 0 is sur- condition is that eo must be much smaller than kT.
rounded by a surface S. 5The reader may wish to skip this mathematical treatment, and resume the argument at Eq. 7.42.
278 Chapter 7 Solutions of Electrolytes

any point on the surface the field strength is E, and for an element of area dS the
scalar product E dS is defined as
EdS=E cos OdS (7.28)
where 6 is the angle between E and the normal to dS. This scalar product is the flux
of the electric field through dS. The element of solid angle dw subtended by dS at
the charge Q is dS(cos 6)/r 2, and the total solid angle dw subtended by the entire
closed surface S at any point within the surface is 47r. These considerations lead to
Gauss's Theorem an expression known as Gauss's theorem:

dS = -- (7.29)
fE fo

In our case the distribution of charge can vary, so that Q is replaced by the
charge density p integrated over the volume enclosed by the surface; that is,

fE dS = _! f pdV (7.30)

It is known from vector analysis that the divergence (div) of a vector E

Divergence of E divE = VE = + + (7.31)


ox ay .Oz

Del Operator is equal to the flux of E from dV per unit volume. Here V is known as the del oper-
ator and when applied to a vector it represents the operation a/ax + 3/y + i9/c3z on
its components in Cartesian coordinates. If Eq. 7.31 is integrated over the system of
charges enclosed by the surface and equated to Eq. 7.30, we have

div EdV=JE.ds= -1-JpdV (7.32)


J

Divergence Theorem The equality on the left is known as the divergence theorem. Since the integrals are
equal for any volume,

divE= - - (7.33)
0

We now relate E to the potential 0. From Eq. 7.4 the electric field strength can be
represented as the gradient of an electric potential and in vector notation
E = grad ct. = V4 (7.34)
Substitution into Eq. 7.33 gives

div E= div grad 4=e (7.35)


0

The operator div grad is often written as V2 and read as "del squared"; it is known
as the Laplacian operator Equation 7.35 can thus be written as

V20 = - (7.36)
Poisson Equation
0
7.4 Strong Electrolytes 279

This is the Poisson equation; the Laplacian operator V2 in it is given by

2 .)2
Laplacian or V --- -----
+
2 + iQ 2
Del-Squared Operator ax'-

For a spherically symmetrical field, such as that produced by an isolated charge, the
Laplacian operator may he written in terms of spherical polar coordinates
The Laplacian operator applied to Eq. 736 then gives

_L _iL (7.38)
r'- iJr itr )
because no dependency on the angles 0 and 0 is found.
The preceding equations are for a vacuum, which has a permittivity e equal to
2
8.854 X It) c'2 i - in . If the potential is in a medium having, a relative per-
mitlivity, or dielectric constant. of e, the absolute permittivity is E1E and must
replace in the preceding equations.
Equation 7.38 is the needed relation to determine p and b in conjLmnction with
Eq. 7.27. Insertion of the expression for p gives

I a /
-TjI = Kb (7.39)
r i \ iir J

where the quantity K (not to he confused with electrolytic conductivity) is given by

=
K Y N (7,40)
E0Eknl '-T

The general solution of Eq. 7.39 is

Ae'
(7.41)

where A and 8 are constants. Since c/ must approach zero as r becomes very large.
and e`/j- becomes very large as r becomes large. the multiplying factor II must be
tern, and Eq. 7.41 thus becomes

Ae'
(7.42)

The constant A can he evaluated by considering the situation when the soluLiun is
infinitely dilute, when the central ion can he considered to he isolated. In such a
,solution /Vz; approaches zero, so that K also approaches j.en and the potential
is therefore

= (7.43)

In these circumstances, however, the potential at a distance r will simply he the po-
tential due to the ion itself, since there is no interference by the atmosphere. The
potential at a distance r due to an ion of charge

(7.44)
41TE?'
280 Chapter 7 Solutions of Electrolytes

Equating these two expressions for 0 leads to

A = (7.45)
47TEE
and insertion of this in Eq. 7.42 gives
= Z(c

4 (7.46)
41r0f r

If K IS sufficiently small (i.e.. if the solution is sufficiently dilute), the exponen-


tial is approximately I - r. and Eq. 7.46 therefore becomes
= ,.e - .(' K
d (7.47)
47reer 47r
The lust term on the right-hand side of this equation is the potential at it distance r.
due to the central ion itself. The second term is therefore the potential produced by
the ionic atmosphere:

ZI?K
& = - (7.48)
47Tf

There is now no dependence on r: this potential is therefore uniform and exists at


the central ion. If the ionic atmosphere were replaced by it charge - z situated at a
distance I/K from the central ion, the effect due to it at the central ion would he ex-
actly the same as that produced by the ionic atmosphere. The distance I/K IS
therefore referred to as the thickness of the ionic atmosphere.
The quantity K IS given by Eq. 7.40: the thickness of the ionic atmosphere is thus

Thickness of the Ionic 1 = / EEkT \I/2


(7.49)
Atmosphere K

Instead of the number of ions N, per unit volume, it is usually more convenient to
use the concentration e,: Ni = eL. where L is the Avogadro constant, and thus

(k3T \112
- ( (7 5()
e
2 (.;,7 L)
K

This equation allows values of the thickness of the ionic atmosphere to be estimated.
Table 7.1 shows values of I/K for various types of electrolytes at different mo-
lar concentrations in aqueous solution. We see from Eq. 7.50 that the thickness of
the ionic atmosphere is inversely proportional to the square root of the concentra-
tion: the atmosphere moves further from the central ion as the solution is diluted.

TABLE 7.1 Thickness of Ionic Atmospheres in Water at 25 'C


Molar Concentration
Type of Electrolyte 0.10 M 0.01 M 0.001 M

Urii-univalent 0.962 mu 3.03 nm 9.62 am


Uni-hivaleni and hi-univaleni 0.556 nni 1.76 am 5,53 am
Bi-bivalent 0.481 am 1-52 am 4.81 am
tJni-tervalent and ter-univalent 0.392 nm 1.24 nm 3.92 nm
7.4 Strong Electrolytes 281

EXAMPLE 7.3 Estimate the thickness of the ionic atmosphere for a solution
of (a) 0.01 M NaCI and (h) 0.001 M ZnCL, both in water at 25 C, with = 78.

Solution

a. The summation in Eq. 7.50 is

0.01 >( l + 0.01 X l = 0.02 niol din

Insertion of the appropriate values into Eq. 7.50 then gives


2) 1112
8.854 X 10 ' -(C N m X 78 X 1.381 X In 1 (J K 'X 298.15(K)
I92((2
K
[ (
l.02 >< 10 X 0.02(niol din ) X 10(dni1 in ) X 6.022 X 102 (rnol )

= 3.03 X it) (I N 1 in)112


in 2 2
Since J = kg s and N = kg m s or J = Nm. the units are metres; the
thickness is thus 3.03 nm (nanometres).

h. The summation is now


2
= 0.001 X 22 + 0.002 X 1 = ().0)6 mol dm

and the result is

= 5.53 X 10 ' ni = 5.53 nm

Mechanism of Conductivity
The mathematical treatment of conductivity is somewhat complicated atid only a
very general account u the main ideas can he given here.
i'he effect of the ionic atmosphere is to exert a drag on the movement of a

0 given ion. If the ion is stationary, the aliiiiisphere is arranged symmetrically about
it and does not tend to move it in either direction (see Fit Lire 7.6a). However, if
a potential that tends to move the ion to the right is applied, the atmosphere will
decay to some extent on the left of the ion and build up more on the right (Fig-
nrc 7.6h). Since it takes lime for these relaxation processes to occur, there will
be an excess of ionic attuosphere to the left of the ion (i.e.. behind it) and a
deficit to the right (in front of it. This asymmetry of the atmosphere will have
the eflct of dragging the central ion hack. This is the relaxation or asymmetry
effect.
There is a second reason why the existence of the ionic atmosphere iiiipedcs
the motion of an ion. Ions are attracted to solvent molecules mainly byion-dipole
FIGURE 7.6
forccs therefore. when they move. they drag solvent along with them. The ionic at-
(a) A stationary central positive ion
with a spherically symmetrical ion mosphere. having a charge opposite to that of the central ion, moves in the opposite
atmosphere. (b) A positive ion ilircetion to it and therefore drags solvent in the opposite direction. This means that
moving to the right. The ion the central ion has to travel upstream, and it thcrelore travels more slowly than if
atmosphere behind it is relaxing, there were no effect of this kind. This is the eleetrophoretic effect.
while that in front is building up.
Dehye and Hiickel carried through a theoretical treatment that led them to ex-
The distribution of negative ions
around the positive ion is now pressions for the forces exerted upon the central inn by the relaxation effect. They
asymmetric. sLipposed the ions to travel through the solution in straight lines, neglecting the
282 Chapter 7 Solutions of Electrolytes

Lars Onsager received the zigzag Brownian IllOtiOfl brought about by the collisions of surrounding solvent
Nobel Prize in chemistry in molecules This theory was improved in 1926 by the Norwegian-American physical
1968 for his discovery of the chemist Lars Onsager (1903-1976), who took Brownian motion into account and
Onsager reciprocal relations, vlise expression for the relaxation fime was/ I

mathematical equations that


are fundamental for the - C
7 SI
thermodynamic theory of - 24k1 T
irreversible processes.
where V' is the applied potential gradient and ii' is a number whose magnitude de-
Relaxation Force pends on the type of electrolyte, for a uni-univalent electrolyte. u' is 2 - =
0.586.
Electrophoretic Force The ele(u'op/u)rezie/oree/ was given by Dehyc and Hckel as

e.K
.1; = ----K, I" (7.52)
6 in

where K, is the coefficient of frictional resistance of the central ion with reference
to the solvent and i is the viscosity of the medium. The viscosity enters into this
expression because we are concerned with the illotion of the ion past the solvent
molecules, which depends on the viscosity of the solvent.
The final expression obtained oii this basis for the molar conductivity A of an
electrolyte solution of concentration c is

Debye-Hckel-Onsager A = A - (P -I- QA)\/c (7.53)


Equation
where I' and Q can he expressed in terms of various constants and properties of the
system. For the particular case of a sv,nn,etrieaf electrolyte I i.e.. one for which the
two ions have equal and opposite signs (c = z P and Q are given by

eP e
(7.54)
3in ( EOfkjjT)

and
0
E 135 / 'e2L P2
(7.55)
133 Silver nitrate 24rEek1 T t 7rEk1)
r. 131
Equation 7.53 is known as the I)ebye-Huckel-Onsager equation. It is based on the
>, 129 assumption of complete dissociation. With weak electrolytes such as acetic acid the
>
127 ions are sufficiently far apart that the relaxation and electrophoretie effects are negli-
chloride gible; the Arrhenius-Ostwald treatment (Section 7.3) then applies. For electrolytes of
c 125
0
C.) intermediate strength a combination of the Dehye-H iickel-Onsager and Arrhenius-
123
0
Ostwald theories must he used,,
, 11
A number of experimental tests have been made of the Debye-Hckel-Onsager
0 0.02 0.04 0.06
equation. Equation 7.53 can be tested by seeing whether a plot of A against Ve is lin-
ear and whether the slopes and intercepts of the lines are consistent with Eqs. 7.54
FIGURE 7.7 and 7.55. An example of such a plot is shown in Figure 7.7. For aqueous .solutions of
Plots of molar conductivity A uni-univalent electrolytes, the theoretical equation is found to he obeyed very satis-
against the square root of the factorily up to a concentration of about 2 X 10 3 mol dm-'-. at higher concentrations.
concentration of the solution.
The plots provide a test of the deviations are found. The corresponding equations for other types of electrolytes in
Debye-Hckel-Onsager equation, water are also obeyed satisfactorily at very low concentrations. but deviations are
Eq. 7.53. found at lower concentrations than with um-univalent electrolYtes.
7.4 Strong Electrolytes 283

Various attempts have been made to improve the Debye-Huckel treatment, One
of these attempts involved taking ion association into account, a matter that is dealt
with in the next subsection. Another attempt, made by Gronwall, La Mer, and co-
workers in 1928, was based on including higher terms in the expansions of the
exponentials in Eq. 7.24. This led to improvement for ions of higher valence, but
unfortunately the resulting equations are complicated and difficult to apply to the
experimental results.
Another weakness of the Debye-Huckel theory is that the ions are treated as
point charges; no allowance is made for the fact they occupy a finite volume and
Manfred Eigen shared the 1967 cannot come close to each other. This difficulty was dealt with in 1952 by E. Wicke
Nobel Prize in chemistry for his and the German chemist Manfred Eigen (b. 1927) in a manner similar to that of in-
development of relaxation troducing the excluded volume b into the van der Waals equation for a real gas
spectrometry. (Section 1.13).

Ion Association
A common type of deviation from the Debye-Hiickel-Onsager conductivity equation
is for the negative slopes of the A versus c plots to be greater than predicted by the
theory; that is. the experimental conductivities are lower than the theory predicts.
This result, and also certain anomalies found with ion activities (Section 7.10), led
to the suggestion that the assumption of complete dissociation sometimes needs
qualification for strong electrolytes. In the case of an electrolyte such as ZnSO4
there are strong electrostatic attractions between the Zn2 and SO_t ions, and it is
possible for pairs of ions to become associated in solution. In this example, in which
the ions have equal charges, the ion pairs have no net charge and therefore make no
contribution to the conductivity, thus accounting for the anomalously low conduc-
tivity. It is important to distinguish clearly between ion association and bond
formation; the latter is more permanent, whereas in ion association there is a con-
stant interchange between the various ions in the solution.
The term ion association was first employed by Niels Bjerrurn, who in 1926
developed a theory of it. The basis of his theory is that ions of opposite signs sepa-
rated in solution by a distance r* form an associated ion pair held together by
coulombic forces. The calculation of Bjerrum's distance r* is based on the Boltz-
mann equations (Eqs. 7.21 and 7.22) and the final result is that, with charge
numbers z and z.

= I (7.56)
8irkT
For the case of a uni-univalent electrolyte (z, Z, = 1) the value of r' at 25 C,
with = 78.3, is equal to 3.58 X i- m = 0.358 nm. The expression for the elec-
trostatic potential energy of interaction between two univalent ions separated by a
distance r is
e
E= (7.57)

Substitution of r* for r (Eq. 7.56) into this expression leads to


E* = 2k8T (7.58)
The electrostatic potential energy at this distance r* is thus four times the mean ki-
netic energy per degree of freedom. The energy at this distance is therefore sufficient
284 Chapter 7 Solutions of Electrolytes

for there to be significant ion association, which will he dynamic in character in that
15 there will be a rapid exchange with the surrounding ions. At distances smaller than
0.358 nm the probability of ionic association increases rapidly. Therefore, to a good
approximation one can say that if the ions are closer than 0.358 nm. they can he con-
10 sidered to be "undissociated' and to make no contribution to the conductivity.
However, in aqueous solution univalent ions call rarely approach one another
as closely as 0.358 nni, and ion association is therefore of little importance for such
ions. In solvents having lower dielectric constants, however, there can he substan-
tial ion-pair formation even for uni-univalent electrolytes. This is strikingly
illustrated by results obtained in 1923 by the American electrochemists Raymond
0 M. Fuoss and Charles A. Kraus. For teiraisoamylanimonium nitrate in a series of
solvents having dielectric constants ranging from 2.2 to 78.6, they made measure-
0 0.5 tO 15 ments of conductances, and from the results they calculated the association
log10 constants K, for ion association. As shown in Figure 7.8, the formation of ion pairs
was negligible in solutions of high dielectric constant, but very considerable in so-
FIGURE 7.8
ILitions of low dielectric constant. Dioxane has a dielectric constant of 2.2 at 25 IV.
A double-logarithmic plot of Ka ,
the association constant for ion- and the r* value for a uni-univalent electrolyte is 12.7 nm: ion association is there-
pair formation with tetraisoamyl- fore expected to he important.
ammonium nitrate, against the Much more ion association is found with ions of higher valence. With salts
dielectric constant of the solvent having ions of different valences, such as NaSO4. ion association will lead to the
(data of A. M. Fuoss and C. A.
formation of species such as Na SO . Again there will be a reduction in conduc-
Kraus, 1923).
tivity, since these species will carry less current than the free ions. Bjerrum's theory
has been extended to deal with such unsymmetrical electrolytes and also with the
formation of triple ions.

Conductivity at High Frequencies and Potentials


An important consequence of the existence of the ionic atmosphere is that the con-
ductivity should depend on the frequency if an alternating potential is applied to the
solution. Suppose that the alternating potential is 01 sufficiently high frequency that
the time of oscillation is small compared with the time it takes for the ionic atnios-
phere to relax. There will then not be time for the atmosphere to relax behind the
ion and to form in front of it; the ion will he virtually stationary and its ionic atmos-
phere will remain symmetrical. Therefore, as the frequency of the potential
increases, the relaxation and electrophoretic effects will become less and less im-
portant, and there will be an increase in the molar conductivity.
A related effect is observed if conductivities are measured at very high poten-
tial gradients. For example, if the applied potential is 20 000 V cm ', an ion will
move at a speed of about I ni s and will travel several times the thickness of the
effective ionic atmosphere in the time of relaxation of the atmosphere. Conse-
quently, the moving ion is essentially free from the effect of the ionic atmosphere,
which does not have time to build up around it to any extent. Therefore, at suffi-
ciently high voltages, the relaxation and electrophoretic effects will diminish and
eventually disappear and the molar conductivity will increase. This effect was first
observed experimentally by the German physicisl Max Carl Wien (1866-1938) in
Wien Effect 1927 and is known as the Wien effect. However, molar conductivities of weak
electrolytes at high potentials are anomalously large, and it appears that very high
potentials bring about a dissociation of the molecules into ions. This phenomenon
Dissociation Field Effect is known as the dissociation field effect.
7.5 Independent Migration of Ions 285

7.5 Independent Migration of Ions


In principle the plots of A against concentration (Figure 7.2) can be extrapolated
back to zero concentration to give the A value. In practice this extrapolation can
only satisfactorily be made with strong electrolytes. With weak electrolytes there is
a strong dependence of A on c at low concentrations and therefore the extrapola-
tions do not lead to reliable A values. An indirect method for obtaining A for
weak electrolytes is considered later.
In 1875 Kohlrausch made a number of determinations of the A values and ob-
served that they exhibited certain regularities. Some values are given in Table 7.2 for
corresponding sodium and potassium salts. We see that the difference between the
molar conductivities of a potassium and a sodium salt of the same anion is indepen-
dent of the nature of the anion. Similar results were obtained for a variety of pairs of
salts with common cations or anions, in both aqueous and nonaqueous solvents.
This behavior was explained in terms of Kohlrausch's law of independent
migration of ions. Each ion is assumed to make its own contribution to the molar
conductivity, irrespective of the nature of the other ion with which it is associated.
In other words,

(7.59)

where A and A are the ion conductivities of cation and anion, respectively, at infi-
nite dilution. Thus, for potassium chloride (see Table 7.2)

A(KC1) = A~ + A_ = 149.9 S cm2 moF' (7.60)


and for sodium chloride

A(NaCI) = A a+ +A- = 126.5 S cm2 moF' (7.61)

The difference, 23.4 S cm moF', is thus the difference between the A values for
K and Na:

- Aa+ = 23.4 S cm moF' (7.62)


and this will be the same whatever the nature of the anion.

TABLE 7.2 Molar Conductivities at Infinite Dilution for Various Sodium


and Potassium Salts In Aqueous Solution at 25 C
A
Electrolyte S cm2 mor1 Electrolyte S cm, mor1 Difference

KC1 149.79 NaCI 126.39 23.4


1(1 150.31 NaT 126.88 23.4
--K2SO4 153.48 --Na2SO4 130.1 23.4

Note that these molar conductivities are what were formerly called "equivalent conductivities." the
Latter term being no longer recommended by IUPAC. The units S cm moF 1 are equivalent to t) -

cm moF1.
286 Chapter 7 Solutions of Electrolytes

Speed = uV Concentration of Ionic Mobilities


positive ions = c,
Charge It is not possible from the A" values alone to determine the individual A and A
+ reaching values however, once one A,'- A' value (such as A,. ) has been determined, the
electrode
in unit time rest can be calculated. The A and AY. values are proportional to the speeds with
+ - =FVc1 which the ions move under standard conditions, and this is the basis of the methods
Unit cube by which the individual conductivities are determined. The mobility ol an ion, a. is
defined as the speed with which the ion moves under a unit potential gradient. Sup-
pose that the potential drop across the opposite faces ol a unit cube is V and that the
concentration of unix ilerit positive ions in the cube is c+ (see Figure 79). If the
Potential drop = V mobility of the positive ions is it the speed of' the ions is it V. All positive ions
within a distance of it V of the negative plate will therefore reach that plate in unit
FIGURE 7.9
A unit cube containing a solution time, and the number of such ions is if 1/(- ,. The charge they carry is Eu V(-, . and
in which the concentration of posi- since this is carried in unit time, the current is Eu .Ve The electrolytic conductiv-
tive ions is c and where there is ity due to the positive ions. K , is thus
a potential drop of V between the
opposite faces. current Fit. 1/c1
K 1 = = = Fit, c, (7.63)
potential drop V

The molar conductivity due to these positive ions is therefore

Eu (7.64)
= -- =

The SI unit of mobility is iii s VV iii = nY V - . but it is more common to


use the unit L:111 V

EXAMPLE 7.4 The mobility of a sodium ion in water at 25 C is 5.19 X

10 4 cm2 V 1 s 1 . Calculate the molar conductivity of the sodium ion.

Solution From Eq. 7.64,

= 96485 As mol X 5.19 X 10 4 cm2 V

= 50.1 A V 1 cm2 inol


I
A V 1 = 11 = S: therefore
0
A .., . = 50.1 S cm2 mol

Transport Numbers
In order to split the A values into the Y and A values for the individual ions we
iiiakc LISC Of a property known variously as the transport number, the transference
iuinthe:; or the migration nuinhe,: It is the /rae!ioiz of the current carried hv each
ion present in solution.
Consider an electrolyte of formula Aa B1,. which ionizes as tullw:

A11B1, aA - bB

The quantity of electricity carried by a mol of the ions A whose charge number
is . is aF: - . and the quantity of electricity crossing a given cross-sectional area
7.6 Transport Numbers 287

in unit time is aPz+u,, where u is the mobility. This quantity is the current car-
ried by the positive ions when 1 mol of the electrolyte is present in the solution.
Similarly, the current carried by b mol of the negative ions is IFZ_U.... The fraction
of the current carried by the positive ions is therefore
aFz+u.4 - az+u+
t+ = (7.65)
aFz,u+ + bFz_u_ az+u+ + bz_u_
This is the transport number of the positive ions. Similarly, the transport number of
the negative ions is
bz. u -
t. = (7.66)
az+u -4- + hz_u..
In order for the solution to be electrically neutral it is necessary that
= -- (7.67)
and Eq. 7.65 and 7.66 then reduce to

Transport Numbers = and t = (7.68)

The individual molar ion conductivities are given by


A(A) = Fzu+ and A(B) = Fz_u_ (7.69)
The molar conductivity of the electrolyte is
A(Aa Bb) = F(az+u+ + bz_u_) (7.70)
It then follows that the transport number t+ is related to the individual ion conduc-
tivities by
A(A )/z
= A(A)/z+ + A(B)/z
together with Eq. 7.67 this gives
- aA(A ) aA(AZ* )
t 1) (7.72)
- aA(A + bA(Bz. _) = A(A,Bh )
Similarly, for t_,
- bA(B) = bA(B)
(7.73)
- aA(A') + bA(B z A(A,Bh )
If t+ and t_ can be measured over a range of concentrations, the values at zero con-
centration, t+ and t, can be obtained by extrapolation. These values allow
A(AaBh) to be split into the individual ion conductivities A(A) and A(B) by
use of Eqs. 7.72 and 7.73.
At first sight it may appear surprising that Faraday's laws, according to which
equivalent quantities of different ions are liberated at the two electrodes, can be
reconciled with the fact that the ions are moving at different speeds toward the
electrodes. Figure 7.10, however, shows how these two facts can be reconciled.
The diagram shows in a very schematic way an electrolysis cell in which there are
equal numbers of positive and negative ions of unit charge. The situation before
288 Chapter 7 Solutions of Electrolytes

Anode Cathode
+
++*4-+++14+414+++-t- 4- ' 444

a.

'1
-
-I.

Discharged
4-4-I- + 444-4414 I -4-4-4-F 4-.-. 4-1-4 I

.4-- - c__I- . ---- - -----


Discharged
b.

I+

+ + -f++ 1-+++ -t -1-++4 +++++ I 4 \ -I/


H Discharged

Discharged

C.

.4 .4

FIGURE 7.10
Schematic representation of the C) S.
movement of ions during an elec-
trolysis experiment, showing that
equivalent amounts are neutral-
ized at the electrodes in spite of 0*000
differences in ionic velocities. Dia-
gram (d) is reproduced from an
original publication by Hittorf. d.

clecliolysis occurs is shown in Figure 7. I Oa. Suppose that the canons only were
able to move, the anions having zero mobility ,- after some motion has occurred.
the situation will he as represented iii I 'igure 7.101). At each electrode two ions i-c-
main unpaired and are dichiarged. two electrons at the same Ilnie traveling in the
outer circuit froni the anode to the cathode. Thus. although univ the cat ions have
moved through the hulk ol the sot Lit OJI. eq u I "a lent aiiou nts have been iii sclia rgcd
at the two electrodes. It is easy to extend this argument to the case in which both
tons are mo inz but at LliltCl'CIlt \peed-.. Thus hgiiie 7.10c '.ho",, the situation
when tile speed of' the cation is three 11111c" that of' the anion, liii lons being dis-
charged at each electrode.
This problem was considered by H ittori'. and Figure 7..l Od is l'L'l)rodUce(l li'oiii
one of his l)tlhlications.

Hittorf Method
Two experimental methods have been employed For the deterini nat ion o(' transport
numbers. One of them. developed by the German physicist Johann \Vj helm Hittorf
1824- 1914)
in 1853. involves measuring the chan ges ut concentiat ion in the vicin-
7.6 Transport Numbers 289

Anode Cathode ity of the electrodes. In the second, the moving boundary method, a study is made
of the rate of movement, under the influence of a current, of the boundary between
two solutions. This method is described on p. 290. A third method, involving the
measurement of the electromotive force of certain cells, is considered in Section 8.5.
A very simple type of apparatus used in the Hittorf method is shown in Figure
7.11. The solution to be electrolyzed is placed in the cell, and a small current is
passed between the electrodes for a period of time. Solution then is run out through
the stopcocks, and the samples are analyzed for concentration changes.
The theory of the method is as follows. Suppose that the solution contains the
ions M andA, which are not necessarily univalent but are denoted as such for sim-
plicity. The fraction of the total current carried by the cations is t,, and that carried
by the anions is t_. Thus, when 96 485 C of electricity passes through the solution,
96 485 tC are carried in one direction by t, mol of M ions, and 96 485 t_C are
compartment compartment
carried in the other direction by t_ mol of A ions. At the same time 1 mol of each
FIGURE 7.11 ion is discharged at an electrode.
Simple type of apparatus used for Suppose that, as represented in Figure 7.12, the cell containing the electrolyte
measuring transference numbers is divided into three compartments by hypothetical partitions. One is a compart-
by the Hittorf method.
ment near the cathode, one is near the anode, and the middle compartment is one in
which no concentration change occurs. It is supposed that the two ions are dis-
charged at the electrodes, which are unchanged in the process. The changes that
occur when 96 485 C is passed through the solution are shown in Figure 7.10. The
results are
At the cathode: A loss of I - = t_ mol of M, and a loss of t_ mol of A

(i.e., a loss of t_ mol of the electrolyte MA).


In the middle compartment: No concentration change.
At the anode: Loss of 1 t_ = t
- mol of A; loss of 1+ mol of M; the net loss
is therefore t, mol of MA.
It follows that

amount lost from anode compartment =


(7.74)
amount lost from cathode compartment r_

FIGURE 7.12 Anode Cathode


The concentration changes in ~
three compartments of a conduc-
tivity cell (Hittorf method). Anode Middle Cathode
compartment compartment i compartment

t+ M t, M'
I mol - - 1 mol
of A of Mt
discharged discharged
tA LA

Net: Net:
f1t no concentration i loss of 1 -

oIofA- change 1 = L mol of M


4 t+ mol of M loss of 1_ mol
290 Chapter 7 Solutions of Electrolytes

Then. since t i- i = 1. the values of 14 and I call he calculated From the experi-
mental results. Alternatively.

amount lost lr'oni anode compartment


=1 (7.75)
amount deposited

and

+ Anode amount lost from cathode compartment


(7.76)
amount (leposi ted

Any of these three equations allows 11 and i to be determined.


M11

In the preceding discusston we have assumed that the electrode, a r e inert


M'A
(i.e.. are not attacked as electrolysis proceeds) and that the ions Ni and A are
M1
-- deposited If instead. Iar example. the anode were to pass into solution during dcc-
trolvsis, the concerriralion changes would he correspondingly different: the treatnient
can readily he modified for this and other situatinn.
MA

b'- --- 41
A Moving Boundary Method
b
MA' The moving boundary method was developed in 1886 by the British physicist Sir
Oliver Joseph Lodge (1851-194W and in 1893 by the British physicist Sir William
('ccii Dampier (formerly VheLham) (1 867- 1952).
The method is illustrated in Figure 7. 13, Suppose that it is necessary to inca-
- Cathode sure the transport numbers ol' the ions in the electrol te MA. Two other electrolytes
a. WA and NIA' are selected as "indicators'': each has an ion in common \N ith MA.
and the electrolytes are such that NI' moes more slowly than NI and A' moves
Cathode Anode
I- I -I more slowlthan A . The solution of NIA is placed in the electrolysis tube with the
Ag,AgCI Pt solution ol M'A on one side of' it and that of' MA' on the other: the electrode ill \i'A
electrode
1 [ electrode the anode and that in MA' is the cathode. As the current flows, the boundaries re-
is
HCI
I r-'Y LiC main distinct, since the NI' ions cannot overtake the NI ions at the hon ndarv u,
solution
solution and the A' ions cannot overtake the A ions at boundary h. The slower ions NI'
+ methyl + methyl
orange and A' are known as the following ions. The boundaries a and h move, as shown
orange
Lii. H in Figure 7.13a. to positions a and h'. The distances an ' and W are proportional to
Capillary
lube _____ the ionic velocities, and therefore

M1 , bb ,
and = 1 (7.77)
(1(1 - 1)/i ii -I- /,b'
b.
Alternatively, the movement of only one boundary may he followed. as shown
FIGURE 7.13
by the example in Figure 7.131). II' Q is the quantity of electricity passed through
The determination of transport
numbers by the moving boundary the sstem. 14 Q is the quantity of' electricity curried by the positive ions, and the
method. (a) Schematic diagram amount of' positive ions to which this corresponds is I t Q/I- . If the concentration of
showing the movement of the ions positive ions is c, the amount 1.. Q/F occupies a volume of t Q/Fc. This volume is
and the boundaries. (b) Simple equal to the area of cross section of' the tube A multiplied by the distance an
apparatus for measuring the trans-
through which the boundary moves. Thus it follows that
port number of H', using Li as
the following cation. A clear bound-
ary IS formed between the reced- = (7,78)
ing acid solution and the LiCI solu- b cA
tion and is easily observed if
methyl orange is present. Since an'. Q. F. c. and A are known, the transport number I can be calculated.
7.7 Ion Conductivities 291

7.7 Ion Conductivities


With the use of the transport numbers the A values can be split into A'+ and A'- the
contributions for the individual ions; the transport numbers extrapolated to infinite
dilution are

L1. = !.. and t-= AA0 (7.79)


A0

Once a single A or A value has been determined, a complete set can be calculated
from the available A values. For example, suppose that a transport number study on
NaCl led to k',,, and A 1 -. If A for KCI is known, the value of 1= A(KCI) -
A 1 -] can be obtained; from A. and A(KBr) the value Of A r can be calculated,
and so on. A set of values obtained in this way is given in Table 7.3. Symbols such as
-Ca2 t are used in the case of the polyvalent ions.
An important use for individual ion conductivities is in determining the A
values for weak electrolytes. We have seen (see Figure 7.2) that for a weak elec-
trolyte the extrapolation to zero concentration is rather a long one, with the result
that a reliable figure cannot be obtained. For acetic acid, however, A is A +
AHcoo ; from the values in Table 7.3 the value of A is thus 349.8 + 40.9 =
390.7 S cm2 moF'.

TABLE 7.3 Individual Molar Ionic Conductivities In Water at 25 C

Cation S cm2 moF' Anion S cm2 moF1

349.65 0H 198
Li 38.66 F 55.5
Na 50.08 CF 76.31
73.48 8r 78.1
Rb 77.8 F 76.8
Cs 77.2 CH3 C00 40.9
Ag 61.9 80.0
T1 74.7 CO 69.3
Mg2+ 53.0
.Ca2+ 59.47
I 2+
-Sr 59.4
fBa2+ 63.6
Cu2 56.6
4Zn2+ 52.8
69.7

Again, these molar conductivities are what were formerly called equivalent
conductivities.
To calculate the mobility u, divide the A. by the Faraday constant 96485 C moF: the
units of u are cm2 V s (see Eq. 7.64 and Example 7.4 on p. 311).
292 Chapter 7 Solutions of Electrolytes

An equivalent but alternative procedure avoids the necessity of splitting the A


values into the individual ionic contributions. The molar conductivity A of any
electrolyte MA can be expressed, for example, as

A(MA) = A(MCI) + A(NaA) - A(NaCl) (7.80)


If MCI, NaA, and NaCl are all strong electrolytes, their A values are readily ob-
tained by extrapolation and therefore permit the calculation of A for MA, which
may be a weak electrolyte. The following example illustrates the use of this method
for acetic acid:
A(CH3COOH) = A(HCI) + A(CH3COONa) - A(NaCI) (7.81)
= 426.2 + 91.0 - 126.5 = 390.7 X Scm2 moF' at 25C
This procedure is also useful for a highly insoluble salt, for which direct determina-
tions of A values might be impractical.

Ionic Solvation
The magnitudes of the individual ion conductivities (Table 7.3) are of considerable
interest. There is no simple dependence of the conductivity on the size of the ion. It
might be thought that the smallest ions, such as Li, would be the fastest moving,
but the values for Li 4 , K, Rb, and Cs show that this is by no means true; K
moves faster than either Li or Na. The explanation is that Li +, because of its small
size, becomes strongly attached to about four surrounding water molecules by ion-
dipole and other forces, so that when the current passes, it is Li(H20) 4 and not Li +
that moves. With Na 4 the binding of water molecules is less strong, and with K it is
still weaker. This matter of the solvation of ions is discussed further in Section 7.9.

Mobilities of Hydrogen and Hydroxide Ions


Of particular interest is the very high conductivity of the hydrogen ion. This ion has
an abnormally high conductivity in a number of hydroxylic solvents such as water,
methanol, and ethanol, but it behaves more normally in nonhydroxylic solvents such
as nitrobenzene and liquid ammonia. At first sight the high values might appear to be
due to the small size of the proton. However, there is a powerful electrostatic attrac-
tion between a water molecule and the proton, which because of its small size can
come very close to the water molecule. As a consequence the equilibrium
H +1120
lies very much to the right. In other words, there are very few free protons in water;
the ions exist as H30 ions, which are hydrated by other molecules (see Figure 7.2 1,
p. 299). Thus there remains a difficulty in explaining the very high conductivity and
mobility of the hydrogen ion. By virtue of its size it would be expected to move about
as fast as the Na ion, which in fact is the case in the nonhydroxylic solvents.
In order to explain the high mobilities of hydrogen ions in hydroxylic solvents
such as water, a special mechanism must be invoked. As well as moving through
the solution in the way that other ions do, the H30 ion can also transfer its proton
to a neighboring water molecule:
H30 4 + H-,0 -4 H20 + H30'
7.7 Ion Conductivities 293

The resulting H30 ton can now transfer a proton to another H20 molecule. In
other words, protons, although not free in solution, can be passed from one water
molecule to another. Calculations from the known structure of water show that the
proton must on the average jump a distance of 86 pm from an HO ion to a water
molecule but that as a result the proton moves effectively through a distance of 310
pm. The conductivity by this mechanism therefore will be much greater than by the
normal mechanism. The proton transfer must be accompanied by some rotation of
HO and H20 molecules in order for them to be positioned correctly for the next
proton transfer. Similar types of mechanisms have been proposed for the hydroxylic
solvents. In methyl alcohol, for example, the process is

CH CH, CH CH

O + 0 - 0 + O
H H H H H

Mechanisms of this type hear some resemblance to it mechanism that was sug-
gested in 1805 by the German physicist Christian Johann Dietrich VOfl Grotthuss
(1795-1822) its it general explanation of conductance and are frequently referred to
Grotthuss Mechanisms as Grotthuss mechanisms.
In non-hydroxylic solvents, in which such mechanisms cannot be involved, the
mobilities of hydrogen and hydroxide ions are no longer abnormally large, as would
he espccted.

Ionic Mobilities and Diffusion Coefficients


Ions also migrate in the absence of an electric field, if there is a concentration gradi-
cnt. The migration of a substance when there is a concentration gradient is known
as diffusion, it topic that will be considered in Further detail in Chapter 19. It will
there he seen that the tendency of a substance to move in a concentration gradient is
measured in terms of it difli,sio,j coelfivient 1). In 1888 the German physicist
Walther Herman Nernst (1864-1941) showed that the relationship between the dif-
fUsiOn coefficient and the mobility 11 Of an ion is

k T
D = u (7.82)

where Q is the charge on the ion. This equation is derived in Section 19.2. Since the
molar ionic conductivity A is equal to Eu and the charge Q is equal to F!zVL, this
equation can he written as

Lk13 TA RTA
I) ( 7.83)
= 1.211 =

These relationships show that diffusion experiments, as well as conductivity mea-


surenients, can provide information about mobilities and hence about molar
conductivities. Equations 7.82 and 7.83 relate to a single ionic species hut, in prac-
tice, experiments must involve at least two types of ions, Diffusion experiments
with a uni-univalent electrolyte A ' B . such as NaCl. will he concerned with the
diffusion of the two ions A and B_, The overall diffusion constant D must then
294 Chapter 7 Solutions of Electrolytes

iclate to the individual diffusion einstaiits /), and D . and Nernst showed that
I'01- uni-univalent electrolytes the proper average is

(7.84)

Walden's Rule
An important relationship jnvOlvimg nioLir conductivity was discovered 111 1906 by
Paul Walden (I S63-1957). who made liiLitstIl'CI))Cfli S of, the molar conductivities of
Ietramelhylammoniuiii iodide in various solvents havint (lit dent Viscosities, The
viseusit ofa liquid, which is treated in Inure detail in Section 19.1. is a measure of
the ease with which a liquid flows. Walden noticed that the product of the molar
conduct i vii . . and the v iscOsltv i of the .sol ciii was approximately constant:

= constant

and this is known as Walden's rule. In Section 19,2 we will inCel the .'.s-/UlSleiI1
((/IWIUfl!. Eq. 19.77, according to which the (lilItIsioll ct)elIIcieiit ui a particle is in-
verselv proportional to the viscosity oh the solvent. Since ionic conduct i vii tes and
(lilt usion coellicients are proportional to one another (Eq. 7.3). Waklcui's 'tile is
eas\ to understand.

78 Thermodynamics of Ions
In thermodynamic work it is customary to obtain values of standard enthalpies and
Gibbs enen.!ies ni peeies. These quantities relate to the formation of I mol of the
substance from the elements in their standard states and UsLiallV icier to 25,0 'C. In
the ease of entropies it is common to obtain absolute values based on the third lass
of t herniudy flaiii ICS.
It is a comparatively straightfoi' aid matter to determine standard enthalpies
and Gibbs enercies of formation and absolute entropies for pmt-.s of' ions in solution.
for example, for a sol iii ion of sodium chloride. We cannot. however, early out ex-
periments on ions ol one kind. The con\ ciii ional procedure is to set the value for
H as 1cm. This all(' s complete sets of values to he built up, and one then speaks
ol (ml'e!lnwwI ,owu1ar(I entholp:e\ o/'/n'ouuio!o. of ('(flIY(',lftW?((l Gibbs energiev
O//c)flflcltUm, and of (ofll'efllu)nal ob.voluie enlropiec.
A few such values for ions are included in Appendix I) p. 1019): more com-
plete compilations are given in Table 7.4. This table includes values of enthalpies.
Gibbs energies, and entropies of hydration: the subscript hvd is used to dciii it
these quantities. These values relate the thermodynamic properties of the ions in
water to their propeities in the as phase and are therefore important in leading to
an uiiderstaiidiiiti of the effect ot tile surroundint water molecules. The (1 ihhs en-
ergv of hydration is the change in Gibbs energy when an ion is transferred horn the
gas phase into aqueous solution. Again. the convention is that the values for the hy-
drogen ion H are all lei-o.
Whereas absolute theiniodynainic values for individual ions cannot he thea-
sured dii'ectiv. they can he estimated on the basis of theory. Liiitortuiiately. ov muL, to
the kiree number of interactions involved. the theoretical treatilteilt Of an ion iii
7.8 Thermodynamics of Ions 295

TABLE 7.4 Conventional Standard Enthalpies, Gibbs Energies, and


Entropies of Formation and Hydration of Individual Ions at 25 C
hydH hydG hydS
Ion kJmoI 1 kJmol' kJmoi' kJmoi 1 JK 1 mor' JlC1 mor1
H 0 (1 0 0 0 0
Li -278.7 576.1 -293.7 579.1 14.2 10.0
Na -239.7 685.3 -261.9 679.1 60.2 21.3
-251.0 769.9 -282.4 751.4 102.5 56.9
Rb - - -282.4 774.0 124.3 69.0
Cs - - -282.0 806.3 133.1 72,0
Ag 105.9 615.5 77.1 610.9 77.0 15.5
Mg2+ - - -456.1 274.1 -118.0 -49.0
Ca2+ -543.1 589.1 -553.1 586.6 -55.2 7.5
Sr2+ - - -557.3 732.6 -39.2 14.2
Ba2 -538.5 879.1 -560.7 861.5 - -
A13 - - -481.2 -1346.4 -313.4 -136.8
F - - -276.6 -1524.2 -9.6 -264.0
CF -167.4 -1469.0 -131.4 -1407.1 55.2 -207.1
Br -120.9 -1454.8 -102.9 -1393.3 80.8 -191.6
F -56.1 -1397.0 -51.9 -1346.8 109.2 -168.6
0H -230.1 -1552.3 - - - -
See Appendix D for sources of thermochemical data. Standard enthalpies of formation and standard
Gibbs energies of formation are from a National Bureau of Standards compilation, "Selected Values of
Chemical Thermodynamical Properties," N. B. S. Circular 500 (1952); heats of hydration. Gibbs
energies of hydration, and entropies of hydration are from D.R. Rossinsky, Client. Rev., 65. 467(1965).

aqueous solution is very difficult. The following absolute values are generally
agreed to be approximately correct, for the proton:
Ahyd H'(absolute) = -1090.8 kJ mol'
4hYdS (absolute) = - 131.8 J K 1 mol -'

LhYdG (absolute) = - 1051.4 ki mol -'


Once these values are accepted, absolute values for the other ions can be calculated
from their conventional values.

EXAMPLE 7.5 Calculate the absolute enthalpies of hydration of Li 4 , 1, and


Ca2+on the basis of a value of -1090.8 kJ mol-1for the proton, using the val-
ues listed in Table 7.4.
Solution The enthalpy of hydration of a uni-univalent electrolyte M X is
independent of whether conventional or absolute ionic values are used. Thus, for
HF, if we lower the value of H by 1090.8, we must raise that for F by the
same amount. Thus the absolute value for F will be
-1397.0 + 1090.8 = -306.2 U mol-'
Similarly, the value for Li must be lowered by 1090.8; the absolute value is thus
576.1 - 1090.8 = -514.7 U mol-1
To obtain the value for Ca2 , we may consider a salt like Cu12 (Cu2 + 21).
Since we have raised the value for each F ion by 1090.8, we must lower the
Ca2 ' value by 2 X 1090.8; thus, 589.1 - (2 X 1090.8) = -1592.5 U mol-'.
296 Chapter 7 Solutions of Electrolytes

7.9 Theories of Ions in Solution


Many theoretical treatments of ions in solution, especially in aqucou solution,
have been put forward. Some of the treatments are quantitative, attempts being
made to obtain expressions for the streng C,th of the interactions between ions and the
solvent. Other treatments are qLlalitative, but based on the mathematical treatments.
Another classification of theories of ions in solution is based on whether the solvent
is treated as if it were a continuum, or whether the molecular nature of the solvent
is taken into account. The latter theories are obviously more realistic. but unfortu-
nately they are difficult to develop in any detail. The continuum theories, such as
the electrostriction treatment of Nernst and Drude and the treatment of Max Born,
are much simpler, and in spite of their lack of' i-ealkiii they give a surprisingly use-
ful interpretation of ionic behavior in Solution.

Drude and Nernst's Electrostriction Model


As early as 1894 Paul Drude (1863-1906) and Waliher Nernst published a paper of
great importance on what they called the electrostriction of, solvent molecules by
ions. They pointed out that the strong electrostatic field due to an ion causes neigh-
boring polar solvent molecules, such as water molecules, to become closely packed
and to occupy less volume. In their treatment, as a matter of mathematical conve-
nience, they regarded the solvent as if it were a continuous dielectric. It was noted
in Section 71 that this electrostriction plays an important role in connection with
the mobilities of ions. and in Section 3.4 that it greatly affects the entropies of ions
in solution.
It was commented earlier (Section 7.4) that one of the criticisms leveled by
some chemists at the theory of electrolytic dissociation was that it appeared to take
no account of the role of the solvent. The recognition by Drude and Nernst that ions
bring about electrostriction of ions played an important role in leading to a general
acceptance of the theory.

Born's Model
A simple interpretation of the thermodynamic quantities for ions in solution was
Max Born did important work suggested in 1920 by the physicist Max Born (1882-1970). In Born's model the
in a number of fields. He did solvent is assumed to be a continuous dielectric and the ion a conducting sphere.
pioneering work on quantum Born obtained an expression for the work of charging such it sphere. which is the
mechanics in the 1920s, and Gibbs energy change during the charging process. The total reversible work in
was awarded the 1954 Nobel
transportincr increments until the sphere has a charge of',-,e is, according to Born's
Prize in physics for his work on
model,
the quantum mechanics of
collision processes.
= - (786)
8'n-E0 e I-

This work, being non-PV work. is the electrostatic contribution to the Gibbs energy
(lithe ion:

(7.87)
81TE(E 1'
7.9 Theories of Ions in Solution 297

If the same charging process is carried out in vacuo ( = I), the electrostatic Gibbs
- energy is
0
z2e2
- G(vacuum)
CS (7.88)
8'rr0r
1000
The electrostatic Gibbs energy of hydration is therefore
+
g2 z 2e2
-2000 hydG = G
es - G(vacuum) = --i (7.89)
80r
2 For water at 25 C, = 78, and Eq. 7.89 leads to
3000 C\
=78 Ga3


I hydG 68.6 _
r/nm
Z
j moF' (7.90)
/ 0AI3
Fe
Figure 7.14 shows a plot of absolute hY dG values against z 2 /r. In view of the
-5000 1 simplistic nature of the model, the agreement with the predictions of the treatments
0 50 100 is not unsatisfactory; the theory accounts quite well for the main effects. Note,
z2/(r/nm) however, that the agreement with the theory is worse the higher the charge and the
FIGURE 7.14 smaller the radius. Figure 7.14 also shows the prediction for = 2, which is ap-
A plot of the absolute Gibbs
proximately the dielectric constant of water when it is subjected to an intense field.
energy of hydration of ions
against z2 /r. The solid line is the The line for = 2 is closer to the points for ions of high charge and small radius.
theoretical line for a dielectric The significance of this will be discussed later.
constant of 78; the dashed line The Born equation also leads to a simple interpretation of entropies of hydra-
is for a dielectric constant of 2. tion and of absolute entropies of ions. The thermodynamic relationship between
entropy and Gibbs energy (see Eq. 3.119) is
(aG\
(7.91)
S= ---)
The only quantity in Eq. 7.89 that is temperature dependent is the dielectric con-
stant , so that
2
= z e2 3 In E
-' es = A byd (7.92)
8ir0r 3T

The reason that the theory leads to the same expression for the absolute entropy of
the ion and for its entropy of hydration is that it gives zero entropy for the ion in the
gas phase, since the Gibbs energy expression in Eq. 7.88 contains no temperature-
0 _ dependent terms. For water at 25 C, is approximately 78 and (3 In )/t3T is about
100 Line of -0.0046 K' over a considerable temperature range. Insertion of these values into
.1
0 L
E -200 Ba2 slope Eq. 7.92 leads to
-4.10 2
) -300- o N/ '-' es by dS = -4.10 J K' moF' (7.93)
(I) -400 -
s Mg2X. Fe r/nm
V 00 Ga3
>.
-500 - Figure 7.15 shows a plot of the experimental hydS values, together with a line of
slope -4.10. The line has been drawn through a value of -42 J K' moF' at z2 /r = 0,
-600 -
since it is estimated that there will be a zhY dS value of about -42 J K' moF' due
700 to nonelectrostatic effects. In particular, there is an entropy loss resulting from the
50 100
z 2 /(r/nm) fact that the ion is confined in a small solvent "cage," an effect that is ignored in the
FIGURE 7.15
electrostatic treatment.
Plot of entropy of hydration Again, the simple model of Born accounts satisfactorily for the main effects.
against z 2 /(r/nm). The deviations are primarily for small ions.
298 Chapter 7 Solutions of Electrolytes

More Advanced Theories


80
Various attempts have been made to improve Born's simple treatment while still re-
60 garding the solvent as continuous. One way of doing this is to consider how the
effective dielectric constant of a solvent varies in the neighborhood of an ion. The
. 40 dielectric behavior of a liquid is related to the tendency of tons to orient themselves
U
a) in an electric field. At very high field strengths the molecules become fully aligned
W
i5 20 in the field and there can he no further orientation. The dielectric constant then falls
to a very low value of about 2. an effect that is known as die/een'ic swuralion.
It has been estimated that the effective dielectric constant in the neighborhood
0 0.4 0.8 1.2 1.6
of ions of various types will he a shown in Figure 7.16. For example. for a ki'ric
Distance from ion / nm
(Fe ) ion the dielectric constant has value of' about 1.78 up to it distance of about
FIGURE 7.16 0.3 nm from the center of' the ion. owing to the high field produced by the ion. At
The variation of dielectric constant greater distances, where the field is less, the effective dielectric constant rises to-
of water in the neighborhood of
ward its limiting value oF 78.
Ions
When the Born treatment of Gibbs energies and entropies of hydration of ions
are modified by taking these dielectric saturation effects into account. tue agree-
ment with experiment is considerably better. The experimental values in Figure
7.12 are consistent with a dielectric constant of' somewhat less than the true value
of 78, and this can he understood ill ternis of the reduction in dielectric constant
arising froill dielectric saturation.

Qualitative Treatments
On the basis of these quantitative theories it is possible to arrive at useful qualita-
tive pictures of ions in solution. The main attractive force between an ion and a
surrounding water molecule is the ion-dipole attraction. The way in which a water
molecule is expected to orient itself in the neighborhood of' a positive or negative
a
ion is shown in Figure 7. 17. There are two possible orientations fur a negative ion.
b. and it appears that both may play a role. Quite strong binding can result from ion-
dipole l'orces: the binding energy for if Na ion and if neighboring water molecule,
FIGURE 7.17
for example. is about 80 ki 11101
Orientation of a water molecule.
(a) Close to a positive ion. The smaller the ion, the greater the binding energy between the 100 and a wa-
(b) Close to a negative ion. ter molecule. because the water molecule can approach the ion more closely. Thus.
the binding energy is particularly strong for the Li ion. and calculations suggest
'

that a Li ion in water is surrounded tetrahedrallv by four water molecules on-

0 f:
o'

14~
Lit

FIGURE 7.18 0
/
/
The tetrahedral arrangement of
four water molecules. (a) Around 0
a Li ion. (b) Around a Cl ion
that can probably also have five
or six water molecules around it. a. b.
7.9 Theories of Ions in Solution 299

Ordered structure of water ented as in Figure 7.17a; this is shown in Figure 7.18. Thus we can say that LC
molecules bound tightly to the ion has a hydration number of 4. The four water molecules are sufficiently strongly
attached to the Li ion that during electrolysis they are dragged along with it. For
this reason (Table 7.3) LF has a lower mobility than Na and K; the latter ions,
being larger, have a smaller tendency to drag water molecules. However, the situa-
tion is not so clear-cut as just implied. A moving Li ion drags not only its four
neighboring water molecules but also some of the water molecules that are further
away from it.
Disorder zone Outer zone A very useful treatment of ions in solution has been developed by the Amer-
in which structure- consisting of ican physical chemist Henry Sorg Frank (1902-1990) and his coworkers. They
making forces ordinary water
oppose each other
have concluded that for ions in aqueous solution it is possible to distinguish
three different zones in the neighborhood of the solute molecule, as shown in
FIGURE 7.19 Figure 7.19. In the immediate vicinity of the ion there is a shell of water mole-
Franks concept of the structure of cules that are more or less immobilized by the very high field due to the ion.
water in the neighborhood of an
These water molecules can be described as constituting the inner hydration shell
ion.
and are sometimes described as an iceberg, since their structure has some icelike
characteristics. However, this analogy should not be pressed too far: Ice is less
dense than liquid water, whereas the iceberg around the ion is more compressed
.
P than normal liquid water.
Surrounding this iceberg there is a second region, which Frank and his co-
workers refer to as a region of structure breaking. Here the water molecules are
( oriented more randomly than in ordinary water, where there is considerable order-
ing due to hydrogen bonding. The occurrence of this structure-breaking region
I;'
results from two competing orienting influences on each water molecule. One of
these is the normal structural effect of the neighboring water molecules; the other
is the orienting influence upon the dipole of the spherically symmetrical ionic
field.
FIGURE 7.20 The third region around the ion comprises all the water sufficiently far from the
The structure of the oxoniurn
(H3O) ion.
ion that its effect is not felt.
The behavior of the proton H in water deserves some special discussion. Be-
cause of its small size, the proton attaches itself very strongly to a water molecule,
l-I + H20 ~ H30
Additional The equilibrium for this process, which is strongly exothermic, lies very far to the
water
molecule right. Thus, the hydrogen ion is frequently regarded as existing as the H30 ion,
held by which is known as the oxonium ion. Experimental studies and theoretical calcula-
tions have indicated that the angle between the 0H bonds in H10 is about 115.
forces
The ion is thus almost, but not quite, flat, as shown in Figure 7.20.
However, the state of the hydrogen ion in water is more complicated than im-
-
plied by this description. Because of ion-dipole attractions, other molecules are

V~~
Water molecules heic
held quite closely to the H30 species. The most recent calculations tend to support
the view that three water molecules are held particularly strongly, in the manner
shown in Figure 7.21. These three molecules are held by hydrogen bonds involving
three hydrogen atoms of the H30 species. The hydrated proton may thus be writ-
by hydrogen bonding ten as H30 3H20, or as H90. Alternatively, some think that the species H90
.

has the same kind of structure as the hydrated Li ion (Figure 7.17a), in which the
FIGURE 7.21 proton is surrounded tetrahedrally by four water molecules. Other water molecules
A possible structure for the
hydrated proton. The hydrated will also be held to H90but not so strongly. Thus'in Figure 7.21 there is shown an
H30 ion (e.g., H90) is called additional H20 molecule held by ion-dipole forces but not by hydrogen bonding,
the hydronium ion. and it therefore is not held so strongly as the three other water molecules. IUPAC
300 Chapter 7 Solutions of Electrolytes

has recommended that the term hydronium ion be employed for these hydrated
HO ions, in contrast to oxonium ion for H30 itself.
The situation is hardly clear-cut; we can write the hydrated proton as
H30, H9O, or H IIO. It is satisfactory to write it simply as H, as long as we re-
member that the proton is strongly hydrated.

7.10 r Activity Coefficients


The electrostatic interactions between ions, besides having an important effect on
the conductivities of solutions of strong electrolytes, have an effect on the ther-
modynamic properties of ions. This matter is most conveniently dealt with in
terms of activity coefficients. Several experimental methods are available for de-
termining the activity coefficients of ionssee Sections 7.11 and 8.5 (emf
measurements). First we will apply the Debye-Huckel treatment to ionic activity
coefficients, as was first done in 1922 by the Danish physical chemist Johannes
Nicolaus Brnsted (1879-1947) and the American physical chemist Victor K. La
Mer (1895-1966).

Debye-Huckel Limiting Law


There are various reasons why a solution shows deviations from ideality, and the
matter is quite complicated. Here we are concerned with a relatively simple reason
for non-ideality in an ionic solution, namely the electrostatic interactions between
ions, as interpreted by the Debye-Huckel theory.
If the behavior of a single ion of type i were ideal, its Gibbs energy could be
expressed by the relationship
Gi = G + kT In c, (7.94)
where ci is the concentration of the ion. Note that in this equation we have used k8 ,
the Boltzmann constant, instead of the gas constant R, since we are concerned with
single ions instead of a mole of ions.
To take into account deviations from ideality we write instead, for the Gibbs
energy,
G f = G + knT in c,y, (7.95)
= G+ kT1n c, + kTIn y (7.96)
The additional term kBT In y1 is due to the presence of the ionic atmosphere; y is
the activity coefficient.
We have seen that the work of charging an isolated ion, on the basis of the
Born model, Eq. 7.86, is
=zi 2e
(797
8ire0r
We also need the work of charging the ionic atmosphere, because this is the re-
quired correction to the Gibbs energy; this work is equal to kT in y. If 0 is the
potential due to the ionic atmosphere, the work of transporting a charge dQ to the
ion is 0 dQ. the net work of charging the ionic atmosphere is thus

w dQ (7.98)
=
7.10 Activity Coefficients 301

The potential 4) corresponding to the charge Q is given by Eq. 7.48 as QK/4ITE0E,


where i/ic is the radius of the ionic atmosphere. The work is thus
,e Qic Q2ic 1z1e_ z,e2ic
fo
-

dQ = (799)
= - ~ 4ir0 8ire0 1 8ir0e
Note that Eq. 7.99 can be obtained from Eq. 7.97 by replacing r by 1/K and chang-
ing the sign. The reciprocal 1/ic plays the same role for the atmosphere as does r for
the ion, and the change of sign is required because the net charge on the atmosphere
is opposite to that on the ion.
The expression for the activity coefficient yj is therefore
- z,eic
k8TIn yj = (7.100)
8ir0
22
z1 e K
whence In y = (7.101)
- 8ire0k8T
ze2ic
or 10 10 91 Yi = (7.102)
- 8 )< 2.303ITEOEkBT
Equation 7.40 shows that ic is proportional to the square root of the quantity N1z,
where Ni is the number of ions of the ith type per unit volume and z1 is its valency.
The ionic strength I of a solution is defined as

Ionic Strength I = CjZ1 (7.103)

where c, is the molar concentration of the ions of type i." The ionic strength is pro-
portional to Ij N1z, and the reciprocal of the radius of the ionic atmosphere, ic, is
thus proportional to \/i (see Eq. 7.49). Equation 7.102 may thus be written as

log10 y = 4BV1 (7.104)

where B is a quantity that depends on properties such as E and T. When water is the
solvent at 25 C, the value of B is 0.51 mo[ 112 dm3 .
Experimentally we cannot measure the activity coefficient or indeed any ther-
modynamic property of a simple ion, since at least two types of ions must be
Mean Activity Coefficient present in any solution. To circumvent this difficulty we define a mean activity co-
efficient y in terms of the individual values for y. and 'y_ by the relationship
=
Y_ (7.105)
where v+ and v_ are the numbers of ions of the two kinds produced by the elec-
trolyte. For example, for ZnCl2, VI = 1 and v_ = 2. For a uni-univalent electrolyte

101t is convenient to use common logarithms in this treatment, as the resulting constant for water at
25 C is then easy to remember (see Eq. 7.111).
'For a uni-univalent electrolyte such as NaCl the ionic strength is equal to the molar concentration.
Thus for I M solution c, = 1 and c._ = I, z.1. = 1 and z_ = I; hence

For a I M solution of a uni-bivalent electrolyte such as K2 SO4, c+ = 2, c_ = 1. z = I, and = 2;


hence the ionic strength is -(2 X I + I X 4) = 3M. Similarly, for a I M solution of a uni-trivalent elec-
trolyte such as Na1PO4, the ionic strength is 6 M.
302 Chapter 7 Solutions of Electrolytes

TABLE 7.5 Mean Activity Coefficients of Electrolytes


as a Function of Concentration
m/rnol kg NaCl NaNO3 Na2HPO4

0.001 ti.9o5
0012 0.952 0.05 1 0.84x
0.005 0.92 0920 I
0.0 0.903 0.900 (1.7)7
0.020 0.872 0.6 0.144
0.050 0822 ()iX I 0 053()
0.101) 0.770 0.750 0.450
0.200 0.734 0.701 0.373
ftSO() 0.0)1! (1.0)7 0.200
1(101) 0.057 0.550 0.101
21)1))) 0.661 0.4)41) ( ). I
5.000 03474 0.3

02
v - = p = I) the mean activity coefticieiil is the geometric mean ( y y of the
individual values. lahle 7.5 shows the eIhci oF di Iie,ciiccs in tvpc of ions within a
compound on the mean activiL coel(tcieiil as the concentration changes.
In order It) express y in terms ol the ionic strength, we proceed as follows.
From Eq. 7.105

+ v ) Ioe y = 104 , 0 y i + 102,111 y (7106)

lnscrtioii of the expression II) Eq. 7.104 for log -y. and log y gives

(vi +P ) 101110 y. = lv. Z i. )fi\// (7.107)

For electrical neutrality

V = vt I or VT. = i (7.108)

and therefore

2 1
(r. + x' ) = B\ (7.109)
( I,

lhus

109111 y = - By! = By! (7,110

For aqueous solutions at 25 C.

log10 y = 0.51:. p VI/moi dm

Equation 7.111 is known as the I)ebye-Hckel limiting law (DI-ILL).

that in thk trea)iniii we ue the pohinve sign tin the \a)enc\ II) iwntive iois tag.. H -
7.10 Activity Coefficients 303

Experimental curve, Deviations from the Debye- Hckel Limiting Law


and prediction
of Eq. 7.115 Experimentally the DHLL is found to apply satisfactorily only at extremely low
concentrations 13 ; at higher concentrations there are very significant deviations, as
o shown schematically in Figure 7.22. Equation 7.110 predicts that a plot of log10 'y
against \/1will have a negative slope, the magnitude of the slope being z,Iz_IB.
Prediction of The results with a number of electrolytes have shown that at very low I values log10
- b Eq. 7.114 'y does indeed fall linearly with I, with the correct slope; for this reason Eq. 7.111
is very satisfactory as a limiting law. At higher V' values, however, the value of
a
----- Prediction of log10 / becomes significantly less negative than predicted by the law and, at suffi-
Slope Debye-Hcket ciently high ionic strengths, may actually attain positive values. The significance of
-z+lz-IB limiting law (Eq. 7.111)
this in connection with the solubilities of salts is considered in Section 7.11.
T
VI Various theories have been put forward to explain these deviations from the
DHLL, but here they can be considered only briefly. One factor that has been ne-
FIGURE 7.22
Variation of log10 y with the glected in the development of the equations is the fact that the ions occupy space;
square root of the ionic strength. thus far they have been treated as point charges with no restrictions on how closely
they can come together. If the theory is modified in such a way that the centers of
the ions cannot approach one another closer than the distance a, the expression for
the activity coefficient of an individual ion (compare Eq. 7.101) becomes
22
ze K
lny1 = - 87rE0 .
kB T I + Ka
(7.112)

Since K is proportional to \/1, this equation can be written as


zB\Il
10910 Y = (7.113)
+aB'V'i
(compare Eq. 7.104), where B is the same constant as used previously and B' is a
new constant. The corresponding equation for the mean activity coefficient y is
z,IzlBV'
10910y= (7,114)
1 +ag-v
(compare Eq. 7.110). Equation 7.114 leads to the curve shown as b in Figure 7.22;
however, it cannot explain the positive values of log10 y that are obtained at high
ionic strengths.
In order to explain these positive values it is necessary to take account of the
orientation of solvent molecules by the ionic atmosphere. This was considered by
Hckel, who showed that it gives a term linear in Fin the expression for log10 'y:

log10 y BzIzl + I (7.115)


= + aB'Vii
where C is a constant. At sufficiently high ionic strengths the last term (CI) pre-
dominates, and log10 y is approximately linear in I, as found experimentally. The
CI term is often known as the "salting-out" term since, as seen in Section 7, 11, it
accounts for the lowered solubilities of salts at high ionic strengths.

13
For example, for a uni-univalent electrolyte such as KCI, the law is obeyed satisfactorily up to a con-
centration of about 0.01 M; for other types of electrolytes, deviations appear at even lower concentrations.
304 Chapter 7 Solutions of Electrolytes

7.11 Ionic Equilibria


Equilibrium is usually established very rapidly between ionic species in solution. In
this section an account will be given of some of the more important ways in which
activity coefficients can be determined when equilibria are established in solution.
Only some special topics are presented here; for a detailed account of how calcula-
tions can be made for fairly complicated systems, the reader is referred to the book
by J. N. Butler in Suggested Reading (p. 304).

Activity Coefficients from Equilibrium


Constant Measurements
Equilibrium constant determinations can provide values of activity coefficients. The
procedure may be illustrated with reference to the dissociation of acetic acid,
CH3COOH H + CH3COO
The practical equilibrium constant is
[W][CH3COO]
Ka = (7.116)
CH3COOHJ YU

where y+ and y_ are the activity coefficients of the ions and y is that of the undis-
sociated acid. In reasonably dilute solution the undissociated acid will behave
ideally (y = 1), but y and y_ may be significantly different from unity because
of the electrostatic interactions. Replacement of y,'y_ by 'y, and taking logarithms
of Eq. 7.116 leads t014
([W][cH1coo]
log,0 [CH = log,0 K - 2 log,0 'y (7.117)
1COOH] )
The left-hand side can be written as
ca
o = log10 K - 2 log,0 y (7.118)
logl ( - a)
1
_
where c is the concentration and a is the degree of dissociation, which can be deter-
mined from conductivity measurements. (See p. 274.) Values of the left-hand side
of this equation can therefore be calculated for a variety of concentrations, and
these values are equal to log 10 K - 2 log 10 y.
If the Debye-HUckel limiting law applies, log,0 y is given by Eq. 7.111. If the
solution contains only acetic acid, the ionic strength I is given by
1= 1[(ca)(l) + (Ca)( 1)2] = ca (7.119)
If other ions are present, their contributions must be added. It is convenient to plot
the left-hand side of Eq. 7.118 against the ionic strength, as shown schematically in
FIGURE 7.23 Figure 7.23. If there were exact agreement with the DHLL, the points would lie on
A schematic plot against V' of
1og10[ca2/(1 - a)], where a, the
degree of dissociation, may be
obtained from conductivity mea- 14
In Eqs. 7.117 and 7.118 we again use the superscript u to indicate the numerical value of the quantity.
surements. K is the thermodynamic equilibrium constant, i.e., the dimensionless form of the practical constant K0.
7.11 Ionic Equilibria 305

it line of slope 2B (i.e.. 1.02 for water at 25 C). Extrapolation to icro ionic strength
gives the value of K. Once Wis value has been obtained, subtraction of log1) K0
horn the experimental curve at any ionic strength gives - Iug y, and y+ can
therefore he calculated at that ionic strength. The individual y and y values can-
not, of course, be determined experimentally.

Solubility Products
When we write sol uhi lity products (Section 4.4), we often ignore activity coefti-
dents; the solubility products are expressed as products of concentrations instead
of activities. The solubility product br silver chloride should more accurately he
written as

K. = IAg'IICI 1-y, y- (7.120)

where y. and y are the activity coefficients of Ag and Cl respectively. The


product y y is equal to y where y is the mean activity coefficient, and
therefore

K=lAgUCliy (7.121)

One matter of interest that can he understood in terms of this equation is the
effect of inert electrolytes on soluhilities. An inert electrolyte is one that does
not contain a common ion (Ag or Cl in this instance) and also does not con-
lain any ion that will complicate the situation by forming a precipitate with
either the Ag or the Cl ions. In other words, the added inert electrolyte does
not bring about a chemical effect ,- its influence arises only because of its ionic
sti-ength.
The influence of' ionic strength I on the activity coefficient of an ion is give!]
according to the DULL by the equation

log114 't = ( 7.122)

Figure 7.22 shows that this equation satisfactorily accounts for the drop in y that
occurs at very low ionic strengths but that considerable deviations occur at higher
ones; the value of log11 y becomes positive (i.e., y is greater than unity) at suffi-
ciently high values of' /.
It follows as a result of this behavior that there are two qualitatively different
ionic-strength etiects on soluhilitics, one arising at low I values when the y falls
with increasing 1 and the other being found when y increases with increasing I.
Thus, at low ionic strengths the product I Ag [(Cl I will increase with increasing I.
because the product [Ag [[('I jy2 remains constant and y, decreases. Under these
Salting In conditions, added salt increases soluhility. and we speak of salting in.
At higher ionic strengths, however, y, rises as I increases, and I Ag [(Cl I
therefore diminishes. Thus there is a decrease in solubility, and we speak of salting
Salting Out out. Of particular interest are the salting-in and salting-out effects found with pro-
tein molecules, a matter of considerable practical importance since proteins Lire
conveniently classified in terms of their solubility behavior.
306 Chapter 7 Solutions of Electrolytes

X 10-11
EXAMPLE 7.6 The solubility product of BaSO4 is 9.2 m012
SO4
Calculate the mean activity coefficient of the Ba2 and ions in a solution
that is 0.05 M in KNO3 and 0.05 M in KC1, assuming the Debye-HUckel limiting
law to apply. What is the solubility of BaSO4 in that solution, and in pure water?
Solution The ionic strength of the solution is
I = f(0.05 + 0.05 + 0.05 + 0.05) = 0.1 M
According to the DHLL,

10910 7+ = _22 X 0.51VI


= 0.645
7+ = 0.226
If the solubility in the solution is s,
= sy = s2 X (0.226)2
9.2 X 10_li = s2 X 0.05126
Therefore
s4.24X l0M
In pure water the activity coefficients are taken to be unity so that the solubility
product is simply the square of the solubility. The solubility in pure water is
therefore
(9.2 X 10 1 ' m012 dm) 9.6 x 10-6 mol dm
We note that the solubility is higher in the salt solution than in water ("salting
in"), because the activity coefficients have been lowered.

We saw earlier that measurements of equilibrium constants over a range of


ionic strengths allow activity coefficients to be obtained. The same can be done
with measurements of solubility. We will outline the method for a sparingly soluble
uni-univalent salt AB, for which the solubility equilibrium is
AB(s) A + B

nj The solubility product is


0
-I- 2
- ------ -Iog 10 7
Thus
A5 = aA+aB = [A ][13 1 (7.123)

- log10 K log1o([A1[B1Y' = log10 K - 2 log10 y (7.124)


For a solution in which no common ions are present, the solubility is
s = [A] = [B} (7.125)
FIGURE 7.24 and therefore
A schematic plot of log10 ?
against the square root of the = log K - 2 log10 y (7.126)
ionic strength, showing how K1
and are obtained. log10 ? = -- log10 K - log10 y (7.127)
7.12 Ionization of Water 307

Insofar as the DHLL is obeyed, a plot of log10 s' against \/!/moi dm 3 will there-
fore be a straight line of slope B = 0.51 in water at 25 C.
Figure 7.24 shows the type of plot that is obtained in this way. At sufficiently
low ionic strengths the points lie on a line of slope B. Extrapolation to zero ionic
strength, where log10 = 0, therefore gives -- log10 K, from which the true
thermodynamic solubility product K is obtained. The value of log y is then given
by the difference between - log10 K and log10 s" at that ionic strength, as shown in
the figure.

712 Ionization of Water


Although water itself ionizes only to a small extent, the fact that it does so at all
plays a vital role in many aspects of physical chemistry. The equilibrium constant
for the ionic dissociation
HOH+OH-
is
[H][OHi
=K (7.128)
[ H201
However, because the concentration of water hardly varies from solution to solu-
tion, it is usual to incorporate this concentration into the equilibrium constant and
Ionic Product to work in terms of the ionic product of water, defined as
K. = [H][OHi (7.129)

At 25 C the value of the ionic product is almost exactly 10_14 mol2 dm.
In chemically pure water the concentrations of hydrogen and hydroxyl ions are
equal and at 25 C are 10-7 mol din 3. A solution in which the hydrogen ion con-
centration is greater than this is said to be acidic; one with a lower concentration is
said to be basic or alkaline.
Acids and bases have been defined in various ways. According to Brnsted, an
acid is a substance that can donate a proton, and a base a substance that can accept
one. The ionization of an acid HA can be written as
HA+H2OH1O+A
and that of a base B as
B+H2OBH+OH_
According to Brnsted's definition the species HA, H3O, and BH are acids,
while B, 0H, and A are bases. In the first process the water is acting as a base, in
the second as an acid. Substances like water that can be both acids and bases are
said to be amphoteric. The species A is said to be the conjugate base of the acid
HA, while BH is the conjugate acid of the base B.
The acid dissociation constant of I-IA is defined as

[H30][A]
K= (7130)
[HA]
308 Chapter 7 Solutions of Electrolytes

It is convenient to work with Pja values, defined as the negative common logarithm
of the Ka value. Thus the dissociation constant of acetic acid is 1.8 X 10-5 M, and
the PKa value is 4.75.
Of special interest are polybasic acids, which undergo successive ionizations.
An example is phosphoric acid, for which the ionizations and the corresponding pK
values are
pK1=2.I pK,'7.2 pK,12.3
H1PO4 H2PO 4 HPO PO
The preponderant species in solution at various pH values can easily be deduced if
the pK values are known (see Problem 7.43).
Amphoteric Electrolytes Also of special interest are amphoteric electrolytes, or ampholytes. Glycine,
for example, ionizes as follows:
H1N'CH2COOH HNCH2COO + H _ H2NCH2COO +2W
The fact that the intermediate species is the zwitterion (double ion) H3N'4 CH2C00
rather than H2NCH2COOH can be shown by an argument such as that in Problem 7.42.

7.13 The Donnan Equilibrium


Of considerable importance are the ionic equilibria that are established when two
solutions are separated by a membrane. Complications arise with ions that are too
large to diffuse through the membrane, and the diffusible ions then reach a special
type of equilibrium, known as the Donnan equilibrium. Its theory was first
worked out in 1911 by the British physical chemist Frederick George Donnan
(1870-1956).
Consider first the equilibrium established when all ions can diffuse through
the membrane. Suppose that solutions of sodium chloride, of volume 1 dm3, are
separated by the membrane, as shown in Figure 7.25a. At equilibrium the concen-
trations are [Na]1 and [CF]1 on the left-hand side and [Na]2 and [CF]2 on the
right-hand side. Intuitively, we know that in this case these concentrations must be
all the same, and in thermodynamic terms we can arrive at this conclusion by not-
ing that at equilibrium

AG = GNa + + G1- = 0 (7.131)

where the terms are the Gibbs energy differences across the membrane (e.g., AG is
the change in Gibbs energy in going from left to right). The expressions for the in-
dividual molar ionic Gibbs energy terms are

[Na]2 [CL]2
= RTIn and G1- = RTIn (7,132)
[Nat ], [C] I
Thus

[Na]2 [CF]2
RTln +RTIn =0 (7.133)
INa11 [CF ]1
from which it follows that

[Na]2[CF]2 -
1 (7.134)
[Na]1[C1ii -
7.13 The Donnan Equilibrium 309

Since, for electrical neutrality, [Na4]1 = [CE], and [Na4]1 = [CFI2, the result is
that
[Na]., Na (Na'-12
[Na'], = [Na 4]2 = ICI-11 = [CF]2 (7.135)

[Cr]1
-
Cr
4-
(Cl-]2 In other words, at equilibrium we have equal concentrations of the electrolyte on
each side of the membrane.
This rather trivial case is useful as an introduction to the Donnan equilibrium,
since Eq. 7.131 is still obeyed even if the system contains a nondiffusible ion in
a.
addition to sodium and chloride ions. For example, suppose that we initially have
the situation represented in Figure 7.25b. On the left-hand side, there are sodium
ions and nondiffusible anions P; on the right-hand side there are sodium and
[Na]1 [Na]2 chloride ions. Since there are no chloride ions on the left-hand side, spontaneous
C4 diffusion of chloride ions from right to left will occur. Since there must always be
electrical neutrality on each side of the membrane, an equal number of sodium ions
[P} [Cr]2
C2 must also diffuse from right to left. Figure 7.25c shows the situation at equilibrium;
x mol dm -3 of [Na4] and [CE] have diffused from right to left; the initial concen-
trations on the two sides are c, and c2 .
b. Application of Eq. 7.134 then leads to

(c2 - x)2 = (Cl + x)x (7.136)


[Na'], Nt. [Na12
c1 +x c2 -x whence

C2
Cl (7.137)
+ 2c2
[Cl-] cr [cl-I2 C,

X C2-X

C.

FIGURE 7.25 EXAMPLE 7.7 Suppose that a liter of a solution of sodium pairnitate at con-
(a) Sodium and chloride ions centration c1 = 0.01 M is separated by a membrane from a liter of solution of
separated by a membrane. sodium chloride at concentration c2 = 0.05 M. If the membrane is permeable to
(b) NaP and NaCl- separated sodium and chloride ions, but not to palinitate ions, what are the final concentra-
by a membrane: initial conditions. tions after the Donnan equilibrium has become established?
(c) The final Donnan equilibrium
conditions arising from (b). Solution Suppose that x mol of sodium ions and .r mol of chloride ions cross
the membrane from right to left. The concentrations are then:
Left side: [Na] = (0.10 + x) M; [CI-] = x M.
Right side: [Na 4] = [CE] = (0.05 - x) M
The products of the concentrations of Na and Cl- ions must be the same on the
two sides, and therefore
(0,01 + x)x = (0.05 - x)2
The solution of this quadratic equation is x = 0.023. The final concentrations are
therefore:
Left side: [Na 4] = 0.33 M; [Cr'] = 0.023 M
Right side: [Na4] = [Cl-] = 0.027 Al
It is easy to check this solution by confirming that the products of the concentra-
tions of the diffusible ions are the same on each side of the membrane.
310 Chapter 7 Solutions of Electrolytes

Equations for more complicated Donnan equilibria (e.g., those involving diva-
lent ions) can be worked out using the same principles. Equilibria of the Donnan
type are relevant to many types of biological systems; the theory is particularly im-
portant with reference to the passage of ions across the membranes of nerve fibers.
However, under physiological conditions the significance of the Donnan effect is
Active Transport not easily assessed, because of the complication of active transport, a phenomenon
in which ions are transported against concentration gradients by processes requiring
the expenditure of energy. One straightforward example of the Donnan effect in
which there is little or no active transport is found with the erythrocytes (red blood
cells). Here the concentration of chloride ions within the cells is significantly
smaller than the concentration in the plasma surrounding the cells. This effect can
be attributed to the much higher concentration of protein anions retained within the
erythrocyte; hemoglobin accounts for a third of the dry weight of the cell.
Under certain circumstances, the establishment of the Donnan equilibrium can
lead to other effects, such as changes in p11. Suppose, for example, that an electrolyte
NaP (where P is a large anion) is on one side of a membrane, with pure water on the
other. The Na ions will tend to cross the membrane and, to restore the electrostatic
balance, H ions will cross in the other direction, leaving an excess of 0H ions.
Dissociation of water molecules will occur as required. There will therefore be a low-
ering of pH on the NaP side of the membrane and a raising on the other side.
The Donnan equilibrium is associated with Nernst potentials, as will be consid-
ered in Section 8.3.

KEY EQUATIONS

Definition of molar conductivity A: Definitions of transport numbers:


U+
A t_.=
C U+ + u_ U, + u_
where where u, and u_ are the ion mobilities.
K = electrolytic conductivity; Definition of ionic strength I:
c = concentration.
ci z12
Ostwald 's dilution law:
c(A/A0)2 = K where
- (A/A) ci = concentration of ion of ith type;
where zi = its charge number.
A = molar conductivity at infinite dilution; Debve-I-Ickel limiting law for activity coefficient y:
A/A = degree of dissociation. 2 r
log10 y= z,B\'l
K = equilibrium constant. For the mean activity coefficient y,
Law of independent migration of ions: login )' = -z,jz_IB"/
A= A + A = 0.51z,IzIVI/mol dm 3
where A and A_ are the individual ion conductivities for water at 25 C.
Problems 311

PROBLEMS
base dissociation constant of ammonia, given the following
Faraday's Laws, Molar Conductivity, molar conductivities at these concentrations: A(K) = 73.5
and Weak Electrolytes 11' cm2 mol'A(C1) = 76.4 fl' cm2 mol'; A(NH) =
73.4 fl-' cm2 moF'; A(0H) = 198.6 fF cm2 moF'.
7.1. A constant current was passed through a solution of
cupric sulfate, CuSO4, for I h, and 0.040 g of copper was 7.9. The conductivity of a 0.0312 M solution of a weak
deposited. Calculate the current (atomic weight of Cu = 63.5). base is 1.53 X 10-4 cm'. If the sum of the limiting ionic
7.2. After passage of a constant current for 45 mm, 7.19 conductances for BH and 0H is 237.0 S cm2 moF',
mg of silver (atomic weight = 107.9) was deposited from a what is the value of the base constant Kb?
solution of silver nitrate. Calculate the current. 7.10. The equivalent conductance of KBr solutions as a
7.3. Electrolysis of molten KBr generates bromine gas, function of concentration at 25 C is given in the following
which can be used in industrial bromination processes. How table. By a linear regression analysis of suitable variables,
long will it take to convert a 500.00-kg batch of phenol find the value of A for KBr.
(C6H5OH) to monobromophenol using a current of 20 000 A? c/10 3 M 0.25 0.36 0.50 0.75
7.4. The following are the molar conductivities A of A/S cm2 m01' 150.16 149.87 149.55 149.12
chioroacetic acid in aqueous solution at 25 C and at vari- 1.00 1.60 2.00 5.00 10.00
ous concentrations c: 148.78 148.02 147.64 145.47 143.15
625 312.5 156.3 78.1 7.11. Equation 7.20 is one form of Ostwald's dilution law.
1t14 M Show how it can be linearized (i.e., convert it into a form
A
fl cm2 mo' 53.1 72.4 96.8 127.7 that will allow experimental values of A at various concen-
F trations to be tested by means of a straight-line plot).
39.1 19.6 9,8 Explain how A and K can be obtained from the plot.
164.0 205.8 249.2 Kraus and Callis, J. Amer Chem. Soc., 45, 2624(1923),
Plot A against c. If A = 362 fl cm2 moF', are these obtained the following electrolytic conductivities K for the
values in accord with the Ostwald dilution law? What is the dissociation of tetramethyl tin chloride, (CH3)4SnCI, in
value of the dissociation constant? (See also Problem 7.11.) ethyl alcohol solution at 25.0 C and at various con-
centrations c:
7.5. The electrolytic conductivity of a saturated solution
of silver chloride, AgC1, in pure water at 25 C is 1.26 X
c/10 4 mol dm 1.566 2.600 6.219 10.441
10 IF' cm' higher than that for the water used. Calcu- 0 1 CM-1
K/106 1.788 2.418 4.009 5.336
late the solubility of AgCl in water if the molar ionic con-
ductivities are Ag, 61.9 I1' cm2 moF'; Cl-, By the use of the linear plot you have devised, deter-
76.4 111 cm2 moF'. mine A and K.
*7.6. The electrolytic conductivity of a 0.001 M solution of 7.12. A certain chemical company wishes to dispose of its
Na2SO4 is 2.6 X fl' cm'. If the solution is satu- acetic acid waste into a local river by first diluting it with
rated with CaSO4, the conductivity becomes 7.0 X 10-4 water to meet the regulation that the total acetic acid con-
fl-' cm* Calculate the solubility product for CaSO4 centration cannot exceed 1500 ppm by weight. You are
using the following molar conductivities at these con cen- asked to design a system using conductance to continu-
Il
trations: A(Na ) = 50.1 fl-1 cm2 mol -i ; A 2-Ca2= ously monitor the acid concentration in the water and trig-
595 cm2 mol'. ger an alarm if the 1500 ppm limit is exceeded. What is
7.7. The quantity IJA of a conductance cell (see Eq. 7.8) is the maximum conductance at which the system should
called the cell constant. Find the cell constant for a conduc- trigger an alarm at a constant temperature of 25 C?
tance cell in which the conductance, G, of a 0.100 M KCI (Assume that the cell constant is 1.0 cm' and that the
solution is 0.01178 S at 25 C. The equivalent conductance density of 1500 ppm acetic acid solution is not apprecia-
for 0.100 M KCI at 25 C is 128.96 S cm mol'. If a bly different from that of pure water. The A for acetic
0.0500 M solution of an electrolyte has a measured conduc- acid is 390.7 S cm2 mol' and K. = 1.81 X 10-5 mol
tance of 0.00824 S using this cell, what is the equivalent dm 3 at 25 C. Ignore the conductance of water.)
conductance of the electrolyte? 7.13. How far can the conductivity of water at 25 C be
*7.8. A conductivity cell when standardized with 0.01 M lowered in theory by removing impurities? The A (in S
KCI was found to have a resistance of 189 Cl. With 0.01 M cm2 mol') for KOH, HCI, and KCI are, respectively,
ammonia solution the resistance was 2460 Cl. Calculate the 274.4, 426.04, and 149.86. K = 1,008 X 10 1 . Compare
312 Chapter 7 Solutions of Electrolytes

your answer to the experimental value of 5.8 X 10-8 the calculations for an initial distance of (a) 1.0 nm to an
cm obtained by Kohlrausch and Heydweiller, Z phys. infinite distance apart and (b) from 1.0 mm to an infinite
Chem. 14, 317(1894). distance apart. (c) In (a) how much work would be required
if the charge is moved to a distance of 0.1 m? The charge
on a proton is 1.6 X l0' C.
Debye-H cket Theory and Transport *7.23. According to Bjerrum's theory of ion association, the
of Electrolytes number of ions of type I present in a spherical shell of
7.14. The radius of the ionic atmosphere (I/K) for a univa- thickness dr and distance r from a central ion is
lent electrolyte is 0.964 nm at a concentration of 0.10 M in dN1 = N, exp(-z jz re2/41rEerk BT) 47rr2 dr
water at 25 C ( = 78). Estimate the radius (a) in water at
a concentration of 0.0001 M and (b) in a solvent of e = 38 where z1 and Z are the charge numbers of the ion of type i
at a concentration of 0.10 M. and of the central ion and e, E, , and kB have their usual
7.15. The molar conductivities of 0.001 M solutions of significance. Plot the exponential in this expression and
potassium chloride, sodium chloride, and potassium sulfate also 41rr2 against r for a uni-univalent electrolyte in water
(+K2s04) are 149.9, 126.5, and 153.3 C1 1 cm2 moF', at 25.0 C (e = 783). Allow r to have values from 0 to
1 nm. Plot also the product of these functions, which is
respectively. Calculate an approximate value for the molar (dN,IN, )dr and is the probability of finding an ion of type i
conductivity of a solution of sodium sulfate of the same at a distance between r and r + dr of the central ion.
concentration. By differentiation, obtain a value r* for which the
7.16. The molar conductivity at 18 C of a 0.0100 M aqueous probability is a minimum, and calculate the value for water
solution of ammonia is 9.6 fl-' CM2 mol '.For NH4CI, A = at 25.0 C. The electrostatic potential is given to a good
129.8 IF1 cm2 moF' and the molar ionic conductivities of approximation by the first term in Eq. 7.47 on p. 280.
0H and Cl- are 174.0 and 65.6 f11 cm2 mo[ 1 , respec- Obtain an expression, in terms of kBT, for the electrostatic
tively. Calculate A for NHI and the degree of ionization in energy between the two univalent ions at this minimum dis-
0.01 M solution. tance, and evaluate this energy at 25 C.
7.17. A solution of LiC1 was electrolyzed in a Hittorf cell.
After a current of 0.79 A had been passed for 2 h, the mass of Thermodynamics of Ions
LiCI in the anode compartment had decreased by 0.793 g.
7.24. The following are some conventional standard
a. Calculate the transport numbers of the Li and Cl- ions.
enthalpies of ions in aqueous solution at 25 C:
b. If A (L1CI) is 115.0 IF' cm2 moF'. what are the molar
ionic conductivities and the ionic mobilities? Ion 1H/kJ moF'
7.18. A solution of cadmium iodide, Cd12, having a molal- H 0
ity of 7.545 X 10 mol kg', was electrolyzed in a Hittorf Na -239.7
cell. The mass of cadmium deposited at the cathode was Ca2 -543.1
0.03462 g. Solution weighing 152.64 g was withdrawn Zn2 -152.3
from the anode compartment and was found to contain CF -167.4
0.3718 g of cadmium iodide. Calculate the transport num- Br -120.9
bers of Cd2 and F.
Calculate the enthalpy of formation in aqueous solu-
7.19. The transport numbers for HCI at infinite dilution are tion of I mol of NaCl, CaCl2, and ZnBr2, assuming com-
estimated to be t = 0.821 and t = 0.179 and the molar plete dissociation.
conductivity is 426.16 1 cm2 mol -'. Calculate the
mobilities of the hydrogen and chloride ions. 7.25. One estimate for the absolute Gibbs energy of hydra-
tion of the H ion in aqueous solution is -1051.4 Id moF'.
7.20. If a potential gradient of 100 V cm' is applied to a On this basis, calculate the absolute Gibbs energies of
0.01 M solution of NaCl, what are the speeds of the Na hydration of the following ions, whose conventional stan-
and Cl- ions? Take the ionic conductivities to be those dard Gibbs energies of hydration are as follows:
listed in Table 7.3 on p. 291.
*7.21. A solution of LiCl at a concentration of 0.01 M is Ion hdG kJ mol-1
contained in a tube having a cross-sectional area of 5 cm2. H 0
Calculate the speeds of the LI' and Cl- ions if a current of Na 679.1
1 A is passed. Use the ion conductivities listed in Table 7.3. Mg 274.1
7.22. What is the work required to separate in vacuum two Al3 -1346.4
particles, one with the charge of the proton, from another CF -1407.1
particle with the same charge of opposite sign? Carry out Br -1393.3
Problems 313

7.26. Calculate the ionic strengths of 0.1 M solutions of of NaCF in aqueous solution at 25 C ( = 78), under the
KNO3, K2SO4, ZnSO4, ZnC12, and K4Fe(CN)6; assume following conditions:
complete dissociation and neglect hydrolysis. a. The electrolyte is present at infinite dilution.
7.27. Calculate the mean activity coefficient 'y for the b. The electrolyte is present at such a concentration that
Ba2 and S02- ions in a saturated solution of BaSO4 the mean activity coefficient is 0.70. The ionic radii are
(K = 9.2 X 10" mol2 dm 6) in 0.2 M K2SO4, assuming 95 pm for Na+ and 181 pm for CF.
the Debye-Hiickel limiting law to apply. 737. If the solubility product of barium sulfate is 9.2 X
7.28. The solubility of AgCl in water at 25 C is 1.274 X 10" m012 dm 6, calculate the solubility of BaSO4 in a
10 mol dm 3. On the assumption that the Debye-HUckel solution that is 0.10 M in NaNO3 and 0.20 M in Zn(NO3)2;
limiting law applies, assume the DHLL to apply.
a. Calculate AG' for the process AgC1(s) - Ag(aq) + 7.38. Silver chloride, AgCI, is found to have a solubility of
CF(aq). 1.561 X 10 M in a solution that is 0.01 M in K2SO4.
b. Calculate the solubility of AgC1 in an 0.005 M solution Assume the DHLL to apply and calculate the solubility in
of K2SO4 . pure water.
7.29. Employ Eq. 7.114 to make plots of log y against v'i 7.39. The enthalpy of neutralization of a strong acid by a
for a uni-univalent electrolyte in water at 25 C, with B = strong base, corresponding to the process
0.51 mol-' dm312 and B' = 0.33 X 1010 mol-1 dm312 m,
and for the following values of the interionic distance a: H(aq) + OH7aq) -* H2O
a = 0. 0.1, 0.2, 0.4, and 0.8 nm is -55.90 Id mol- 1. The enthalpy of neutralization of HCN
7.30. Estimate the change in Gibbs energy AG when 1 mol by NaOH is - 12.13 Id mol '. Make an estimate of the
of K ions (radius 0.133 nm) is transported from aqueous enthalpy of dissociation of HCN.
solution ( = 78) to the lipid environment of a cell mem- 7.40. Make use of the Debye-HUckel limiting law to esti-
brane ( = 4) at 25 C. mate the activity coefficients of the ions in an aqueous
7.31. At 18 C the electrolytic conductivity of a saturated 0.004 M solution of sodium sulfate at 298 K. Estimate also
solution of CaP2 is 3.86 X 10 [F' cm', and that of pure the mean activity coefficient.
water is 1.5 X 106 [Fl cm 1 . The molar ionic conductiv-
ities of ka2+ and F are 51.1 fl-' cm2 mol-' and 47.0
(1! cm2 moF', respectively. Calculate the solubility of
Ionic Equilibria
CaF2 in pure water at 18 C and the solubility product.
7.41. A 0.1 M solution of sodium palmitate,
7.32. What concentrations of the following have the same C,5H31COONa, is separated from a 0.2 M solution of
ionic strength as 0.1 M NaCl? sodium chloride by a membrane that is permeable to Na
CuSO4, Ni(NO3)2, Al2(SO4)3, Na3PO4 and Cl- ions but not to palmitate ions. Calculate the con-
centrations of Na and Cl- ions on the two sides of the
Assume complete dissociation and neglect hydrolysis. membrane after equilibrium has become established. (For a
7.33. The solubility product of PbF2 at 25.0 C is 4.0 X calculation of the Nernst potential, see Problem 8.18.)
10 m013 dm-9. Assuming the Debye-HUckel limiting law 7.42. Consider the ionizations
to apply, calculate the solubility of PbF2 in (a) pure water
and (b) 0.01 M NaE H + H3NCH2COO H3NCH2COOH
7.34. Calculate the solubility of silver acetate in water at H2NCH2COOH + H
25 C, assuming the DHLL to apply; the solubility product
is 4.0 x 10-3mo!2 dm 6. Assume that the following acid dissociation constants apply
to the ionizations:
'7.35. Problem 7.30 was concerned with the Gibbs energy
change when 1 mol of K ions are transported from water H2 + H; Ka = 1.5 x 10 M
to a lipid. Estimate the electrostatic contribution to the -COOH -000 + H; Ka = 4.0 X 10 M
entropy change when this occurs, assuming the dielectric
constant of the lipid to be temperature independent, and the Estimate a value for the equilibrium constant for the
following values for water at 25 C: = 78; a in e/3T process
-0.0046 K'. Suggest a qualitative explanation for the sign
H3NCH3COO H2NCH2COOH
of the value you obtain.
*7.36. Assuming the Born equation (Eq. 7.86) to apply, 7.43. The pK values for the successive ionizations of phos-
make an estimate of the reversible work of charging 1 mol phoric acid are given on p. 308. Which of the four species
314 Chapter 7 Solutions of Electrolytes

is predominant at the following values of the hydrogen or Essay Questions


hydroxide concentration?
7.45. State Faraday's two laws of electrolysis and discuss
a. [H] 0.1 M. their significance in connection with the electrical nature of
b. [H] = 2 x 10M. matter.
c. [WI = 5 x 10 M.
7.46. Discuss the main ideas that lie behind the Debye-
d. [OH-1 = 2 x 10 M.
Hckel theory, as applied to the conductivities of solutions
e. [0H] = 1 M.
of strong electrolytes.
7.44. Two solutions of equal volume are separated by a
7.47. Outline two important methods for determining
membrane which is permeable to K and CF ions but
transport numbers of ions.
not to P ions. The initial concentrations are as shown
below. 7.48. Explain why Li has a lower ionic conductivity than
Na and why the value for H is so much higher than the
[K] = 0.05 M [K] = 0.15 M values for both of these ions.
[Cr] = 0.05 M I [P1 = 0.15 M 7.49. Describe briefly the type of hydration found with the
Calculate the concentrations on each side of the membrane following ions in aqueous solution: Li, Br, H, OW,
after equilibrium has become established. (See Problem 7.50. What modifications to the Debye-HUckel limiting
8.26 in Chapter 8 for the calculation of the Nernst potential law are required to explain the influence of ionic strength
for this system.) on solubilities?

SUGGESTED READING

For general accounts of electrochemistry, see B. E. Conway, ionic Hydration in Chemistry and Bio-
physics, Amsterdam: Elsevier North Holland, 1980.
J. M. G. Barthal, K. Krienke, and W. Kunz, Physical Chem-
B. E. Conway and R. G. Barradas, (Eds.), Chemical
istry of Electrolyte Solutions: Modern Aspects, New
Physics of Ionic Solutions, New York: Wiley, 1966.
York: Springer-Verlag, 1998.
C. W. Davies, Ion Associations, London: Butterworth,
J. O'M. Bockris and A. K. N. Reddy, Modern Electrochem-
1962.
istry. 2 volumes, New York: Plenum Press, 1973. Vol-
D. Eisenberg and W. J, Kauzmann, The Structure and Prop-
ume 1 deals with ionic solutions, Volume 2 with the
erties of Water Oxford: Clarendon Press, 1969.
kinetics of electrode processes.
R. M. Fuoss and F. Accascina, Electrolytic Conductance,
C. M. A. Brett and A. M. 0. Brett, Electrochemistry: Prin-
New York: Interscience, 1959.
ciples, Methods, and Applications, Oxford: University
H. S. Harned and B. B. Owen, The Physical Chemistry of
Press, 1993.
Electrolytic Solutions, New York: Reinhold, 1950 (3rd
S. Glasstone, An introduction to Electrochemistry, New
ed., 1958).
York: Van Nostrand Co., 1942. This old book, long out
H. S. Harned and R. A. Robinson, Multicomponent Elec-
of print but still to be found in libraries, gives an
trolyte Solutions, Oxford and New York: Pergamon
extremely clear account of the subject.
Press, 1968.
G. Korttim, Treatise on Electrochemistry, Amsterdam: Else-
J. L. Kavanau, Water and Solute-Water Interactions, San
vier Publishing Corporation, 1965.
Francisco: Holden-Day, 1964.
J. T. Stock and M. V. Orna (Eds.), Electrochemistry, Past
L. Onsager, "The Motion of Ions: Principles and Concepts,"
and Present, American Chemical Society, Washington,
Science, 166, 1359(1969).
D.C., 1989.
J. Robbins, ions in Solution, Oxford: Clarendon Press,
The following publications deal with special aspects of 1972.
electrolyte solutions. R. A. Robinson and R. H. Stokes, Electrolyte Solutions
(2nd ed.), London: Butterworth, 1959.
N. Butler, Ionic Equilibrium: A Mathematical Approach,
Reading, MA: Addison-Wesley, 1964.

You might also like