70-6861 reed, james frederick,1925 - ohiolink etd center

211
This dissertation has been 70-6861 microfilmed exactly as received REED, James Frederick,1925- DETERMINATION OF THE COMPLEXITY CONSTANTS OF TRANSITION METAL HALIDES AND POLYPHOSPHATES. The Ohio State University, Ph.D., 1969 Chemistry, analytical University Microfilms, Inc., Ann Arbor, Michigan

Upload: khangminh22

Post on 16-Jan-2023

1 views

Category:

Documents


0 download

TRANSCRIPT

This dissertation has been 70-6861microfilmed exactly as received

REED, James Frederick,1925- DETERMINATION OF THE COMPLEXITY CONSTANTS OF TRANSITION METAL HALIDES AND POLYPHOSPHATES.

The Ohio State University, Ph.D., 1969 Chemistry, analytical

University Microfilms, Inc., Ann Arbor, Michigan

DETERMINATION OF THE COMPLEXITY CONSTANTS OF

TRANSITION METAL HALIDES AND POLYPHOSPHATES>

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

By

James Frederick Reed, B.S., M.A.

* * * * *

The Ohio State University 1969

Approved by

ACKNOWLEDGMENTS

To all who have helped make this work possible, I especially wish to thank the following:

My preceptor, Dr. James I. Watters, who was always ready with helpful suggestions as my work continued at The Ohio State University;

the National Science Foundation which gave me the opportunity to study and review chemistry before I returned to graduate school;

Dr. Richard Wynne, of Westinghouse Electric Corporation, who inspired me to continue in the field of analytical chemistry;

Dr. Walter A. Kearney, of the McKeesport Campus, Pennsylvania State University, who gave me the chance to enter the academic field and encouraged me to return to graduate school;

Mrs. Laura Beach, who drew most of the figures and graphs;

Mrs. June Reed, my wife, who typed most of this report.

ii

UVITA

Birth:

Education:

Industrialand

TeachingExperience:

March 3, 1925, New Castle, Pennsylvania

High School, Kingwood, West Virginia, 1942

Westminster College, New Wilmington, Pennsylvania B.C. in Chemistry, 1946

The Ohio State University, Columbus, Ohio,M.A. in Analytical Chemistry, 1948

National Science Foundation Study Grant,Emory University, Atlanta, Georgia, Summer, 1965

Graduate Study at The Ohio State University in Analytical Chemistry, 1966-1969

Westinghouse Electric Corporation, Research Laboratories, Pittsburgh, Pennsylvania} Inorganic Analysis, 1948-1962

Pennsylvania State University, McKeesport Campus; General and Analytical Chemistry, 1962-1966

The Ohio State University, Columbus; Lecturing in Quantitative Analysis, 1966-1969

iii

CONTENTS.

PageACKNOWLEDGMENTS ........................ ............. . . . . . . 11

V I T A ............................................... .. . . . . . . Ill;

T A B L E S ............................................................ vi

ILLUSTRATIONS ...................................................... ix

SYMBOLS................... . . . xi

ChapterI. NATURE OF THE HALIDE COMPLEXES OF COBALT AND RELATED

ELEMENTS ................................................. 1Introduction Historical Review Other Cobalt ComplexesHalide Complexes of Certain Platinum Metals Summary

II. ELECTRICAL AND ION EXCHANGE EXPERIMENTS . ............ 27Electrical Migration Electrophoresis Ion Exchange

III. SPECTROPHOTOMETRY OF COBALT(ll) AND COPPER(ll) INAQUEOUS CHLORIDE SOLUTIONS .............................. 35

IntroductionSpectrophotometry with Common Anions Spectrophotometry in Hydrochloric Acid Spectrophotometry at High Ionic Strength Copper(ll) in Hydrochloric Acid Solt .ion Cobalt in Hydrobromic Acid Solution Summary

IV. COBALT(II) CHLORIDE COMPLEXES IN ORGANIC SOLVENTS . . . . 67Solvent ExtractionProperties of Cobalt(ll) Chloride in AcetoneUse of Spectrophotometry to Determine Stability Constantsby the Slope-Intercept MethodDetermination of Stability Constants by Means of Corresponding Solutions

iv

Effect of Water in the AcetoneSpectra of Cobalt Bromide SolutionsSpectra of Cobalt Iodide SolutionsConductance Experiments in AcetoneSolutions of Cobalt Chloride in the Lower AlcoholsSummary

V. THE DETERMINATION OF HALIDES WITH COBALT(ll) IN ACETONE . 123Introduction Experimental

ProcedurePreparation of Standard Curves

Results and DiscussionComparison of the Bromide and Chloride Absorption CurvesThe Iodide Absorption Curves Interferences Accuracy and Precision Comparison with Volumetric Method Adherence to Beer's Law

SummaryVI. THE COMPLEXES OF TRANSITION METALS WITH POLYPHOSPHATES . . 146

Introduction ExperimentalNotes on Experimental Technique

Nitrogen Atmosphere Ion Exchange Effect of Chloride TemperaturePoints Recorded During Titration Order of Addition of Reagents

Calculation of Metal-Polyphosphate Stability ConstantsSample Calculation of Cu(ll) PyrophosphateDiscussionSummary

REFERENCES.........•............ - ' . 188

v

4

7

9

14

15

30

49

52

53

54

56

60

61

61

64

65

TABLES

u

Absorption Peaks of Cobalt Complexes . . . . . . . . . .

Absorption Peaks of Cobalt Halide Complexes . . . . . .

Absorption Peaks of Chloride Complexes in Acetone . . .

Association Constants of Transition Metal Chlorides . .

Absorption Data for Solvent Complexes of Cobalt(ll) . .

Migration Under Electrophoresis . . . . . . . . . . . .

Absorbance Data for for CoClg in HC1 . . . . . . . .

Absorbance Data for CoCl^ in Perchloric Acid at (i s 12 .

Absorbance Data for C o C ^ in Perchloric Acid at [i = 8.0

Absorbance Data for C o C ^ in Perchloric Acid at (i = 10.0

Absorbance Data for C o C ^ in LiNO^ Solution ■ • • • • •

Absorbance of Copper(II) in Hydrochloric Acid . . . . .

Absorbance Data for CuCl^ for in HCIO^ . . . . . . .

Absorbance Data for K, for CuCl+ in HC10. . . . . . . . .1 4

Absorbance Data for C o B ^ in Perchloric Acid . . . . . .

Stability Constants in Aqueous Systems ......... . . . .

vi

-217. Data for for CoCl^ in Acetone .......... 75

18. Normalized Absorbance Data for CoCl^ at 660 n m. . . . . . . 85

19. Normalized Absorbance Data* for C o C ^ at 670 n m .............. 86

20. Normalized Absorbance Data for CoCl^ at 680 n m . . . . . . . 87

21. Normalized Absorbance Data for CoCl^ at 690 n m . . . . . . . 88

22. Normalized Absorbance Data for CoCl^"" at 674 n m ........... 90

23. Normalized Absorbance Data for CoCl^ at 694 nm ........... 91_224. Absorbance Data for K. of CoBr, in Acetone at 715 nm . . . 954 4

25. Normalized Absorbance Data for CoBrg at 670 n m. . . . . . . 96

26. Normalized Absorbance Data for CoBrg at 680 n m ............ 97

27. Normalized Absorbance Data for GoBr^ at 680 nm . . . . . . 100

28. Normalized Absorbance Data for CoBr^" at 704 n m ........... 101

29. Absorbance Data for Co-I System at 664 nm . . . . . . . . . 106

30. Absorbance Data for Co-I System at 710 nm . . . . . . . . . 107

31. Absorbance Data for Co-I System at 735 nm . . . . . . . . . . 108

32. Absorbance Data for Co-I System at 745 nm . . . . . . . . . 109

33. Absorbance Data for Co-I System at 755 nm 110

34. Comparison of Halide Stability Constants in Acetone with 111Those of Fine . . . . . . . . . . . . . . . . . . . . . . .

35. Summary of Absorbances of Cobalt Halide Complexes in Acetone 112

36. Relative Conductance of Acids in Acetone ................. 113

vii

37. Relative Conductance of Salts in Acetone 115

38. Absorbance Data for CoCl^ in Methanol ..................118

39. Absorbance Data for 0^ f°r CoCl,, in Ethanol at 662 nm . . . 120

AO. Absorbance Data for for CoCl^in Ethanol at 603 nm . . . 120

-2Al. Absorbance Data for for CoCl^ in Ethanol at 689 nm . . 121

A2. 'Summary of Stability Constants of CobaltHalides in Organic S o l v e n t s ................ 122

A3. Effect of Other Anions on the Determination of Halides . . . 141

AA. Summary of Chloride Results . . . . . . . . . . . 142

45. Summary of Bromide Results ............ 143

46. Comparison of CoCl^ Method with AgNO^ Titration . . . . . . 144

47. Comparison of Titrations in Chloride and Nitrate Solutions . 160

48. Titration of Copper Pyrophosphate . . . . . . . . . . . . . 167

49. Titration Data for Polyphosphates . . . . . . . . . . . . . 172

50. Comparison of Results with Those of Other Workers . . . . . 174

51. Metal-Polyphosphate Constants . . . . . . . . . . . . . . . 180

52. Effect of Concentration on Stability Constant Calculationsof Nickel Pyrophosphate 183

53. Effect of Concentration on Stability Constant Calculationsof Cobalt(ll) Triphosphate ........... • . . . . . 184

54. Effect of Concentration on Stability Constant Calculationsof Cobalt(ll) Tetraphosphate..................... 184

viii

ILLUSTRATIONSI)

Figure

1. Electrical Migration Apparatus . ........... . . . . . . .

2. Absorbance of Co(ll) in HNO^ . . . . . . . . . . . . . . .

3. Absorbance of Co(ll) in . . . . . . . . . . . . . .

4. Absorbance of Co(ll) in Concentrated H C 1. . . . . . . . .

5. Absorbance of Co(ll) in 500-700 nm Region . . ^ . . . . .

6. Absorbance of Various Concentrations of Cobalt in 12F HC1 at 694 nm .................... . . . . . . . . . . . . . .

7. Absorbance of 0.002M Co++ in HC1 of Varying Concentrations at 694 n m ..................................... ..

8. Graph for 0,, for CoCl^ in H C 1 ............. ...............

9. Graph for f°r CoCl^ at p. = 1 0 ....................

10. Absorbance of Cu(ll) in HC1 . . . . . . . . . . . . . . ./

11. Absorbance of Co(ll) in HBr ............... . . . . . . .

12. Extraction of Cobalt Chloride with n-Heptanol from 12F HC1

13. Absorbance Spectrum of Co(ll) in Acetone with Added LiCl

14. Absorbance of Co(ll) in Acetone with Added Cl at Selected Wavelengths . . .................... ........... . . . . . .

-215. K. for CoCl. in A c e t o n e ............................ ..4 4

16. Absorbance Curves of Two Corresponding Solutions in Cobalt-Iodide System

17. Series of Corresponding Solutions for Cobalt Chloride . . .

18. Absorbance of Cobalt Chloride in Various Solvents . . . . . 92

19. Absorbance of Co(ll) in Acetone with Added Lithium Bromide . 94

20. Corresponding Solutions for C o B ^ in Acetone at 670 nm . . . 98

21. Absorbance of Co(ll) Iodides in Acetone . . . . . . . . . . 103

22. Corresponding Solutions for Co-I System in Acetone at 735 nm 104

23. Absorbance of Co(ll) + LiCl in Methanol 116

24. Absorbance of Co(ll) + LiCl in Ethanol . ' . ............ 119

25. Determination of Chloride in Acetone at 675 n m .......... 132

26. Determination of Bromide in Acetone at 677 nm . . . . . . . 133

27. Absorbance of CoX^ in Acetone . . . . . . . . . . ......... 135

28. Absorbance of Cobalt Iodide in Acetone at Constant Iodide Concentration 137

4

29. Absorbance of Cobalt Iodide in Acetone at Constant CobaltConcentration........................ 138

30. Effect of Nitrate Ion on the Absorbance of Cobalt Chloride . 140

31. Titration of M(ll) Tetraphosphates in 1:1 R a t i o .............. 154

32. Titration of M(ll) Triphosphates . . . . . . . . . . . . . . 155

33. Titration of M(ll) Pyrophosphates . . . . . . . . . . . . . 156

34. Titration of Various Ratios of Zn and Triphosphate......... 186

x

DEFINITIONS OF SYMBOLSO

percent transmission of radiant energy through medium

ebc = absorbance of radiant energy by a solution

the absorbance coefficient or absorptivity

the light path in centimeters

the concentration in moles per liter

the total or analytical concentration of X in moles per liter

the concentration of X in the indicated form

the activity of H as measured on a pH meter

the activity coefficient

2 2 2the d orbitals, z and x -y , which point directly toward octahedrally bound ligands

the d orbitals, xy, xz, and yz, which point between octahedrally bound ligands

the energy transition between e and t„ orbitalsg . 2g

A/10

the heat of a reaction, usually in kilocalories per mole at 25°C

the entropy of a reaction in entropy units per degree per mole . 9mfl = 10 meter

xi

wavelength of radiant energy

the association constant for the addition of a single ligand in the reaction

oM - + L = ML n-1 n

[MLn]K =-

the overall association constant for the reaction

M + nL « ML [MLn]n

Pn " ------ «n M M *

the disassociation constant of an acid-T+-DTDtL-]

K = a [HL]

the ionic strength in moles per liter

s= mole fraction of X present

xii

CHAPTER I

NATURE OF THE HALIDE COMPLEXES OF COBALT AND RELATED ELEMENTS

Introduction

The objective of this study was to determine the formula and

stability of the blue cobalt chloride complex in concentrated hydrochlo­

ric acid. Because of its blue color and absorption of the relatively

weak red end of the spectrum, most scientists have assumed the complex

to be tetrahedral. The formulas CoC^, CoCl^**, or CoCl^” , have been

assigned without proof to the colored complex by various researchers.

There is a great amount' of literature on the chloride complexes

of cobalt(ll), but much of it is contradictory. Most assume the formula

of the pink hydrated ion to be Co(H2 0 )g++ and its structure to be octa­

hedral. In concentrated hydrochloric acid a blue color develops due to

the formation of a complex assumed to be tetrahedral, CoCl^C^O)^

where n can be any number from 1 to 4; the charge is 2-n. The same blue

color develops in organic solvents containing cobalt and chloride ions,

but the cloride concentration can be very much smaller. Complexes of

cobalt(lll) and chloride are not sufficiently stable in aqueous solution

to prevent the oxidation of chloride to chlorine by cobalt(lll). Partial

replacement of other ligands such as ammonia with chloride can occur.

1

2

Historical Review

When cobalt(ll) forms a halide complex, there is a shift in

the absorbance maximum to a longer wavelength and the absorptivity

increases very greatly. This indicates a definite change in structure

as the hydrated ion is changed to a halide complex. For metals in the

second and third long rows of the periodic chart, there are only small

changes in the absorbance maximum and absorptivity because the basic

structure remains the same, usually octahedral. In the case of cobalt

the absorbance maximum for the octahedral hydrated ion, CoC^O)^**, is

at 514 nanometers, and the molar absorptivity is 4.555 (Jorgensen 1954).-2In nitromethane, where tetrahedral CoCl^ is presumed to form, the

absorbance maximum is at 693 nm and the molar absorptivity is 653 (Gill

and Nyholm 1959). The conversion of only a small fraction of the cobalt

to the tetrahedral chloride complex makes the solution appear blue,

green, or violet.

Cobalt(ll) is chemically more similar to its horizontal than

its vertical neighbors in the periodic chart. All the divalent metal

ions from manganese to zinc form tetrachloro complexes in the solid

state. In aqueous solution only Fe(lll), Co(ll), and Cu(ll) form

tetrahedral chloro complexes. The stabilities of the halide complexes

In organic solvents are intermediate and depend upon the dielectric

constant and donor strength of the solvent (Katzin 1962). Copper(ll)

is unusual in that it forms both square-planar and tetrahedral halide

complexes In glacial acetic acid (Eswien et al. 1967A). In aqueous

solutions the complex appears to be square-planar. Copper, like cobalt,

3

undergoes a large increase in absorptivity when the coordination number

changes from six to four. However, the absorbance maximum moves toward

shorter wavelengths.

Since the spectra of crystalline cobalt tetrachlorides, such as

CSgCoCl^, are very similar to that of cobalt(ll) in concentrated hydro-‘ - 2 chloric acid, many researchers have assumed that the CoCl^ ion occurs

in both. Few have even tried other means of measurement, such as

electrophoresis, chromatography, and electrical conductivity. Some have

measured the magnetic moment, but this evidence is complicated by the

fact that both tetrahedral and octahedral complexes of cobalt(ll) have

three unpaired electrons.

The cobalt(II) ion has a 3d^ electronic structure. Since there

are three unpaired electrons, all d orbitals are occupied. Bonding with

the small highly electonegative atoms, oxygen and nitrogen, is usually

octahedral. The larger chlorine and heavier halogen atoms seem to prefer

to arrange tetrahedrally around a cobalt ion. Magnetic measurements

still show three electrons (Gill and Nyholm 1959), and the bonding may

be sp . Some of the negative charge of the chloride ions is transferred

to the central cobalt ion.-2The electronic transition in CoCl^ is apparently from 3d to

4p, and this is permitted by electronic energy rules (Gill and Nyholm

1959). This transition for the tetrahedral cobalt(ll) ion has been

measured as 14,400 cm”^. The absorbance of the hydrated ion, Co^gO)g++,

is weak because a "forbidden" d-d electronic transition is involved.

However, more energy, 19,500 cm , is required for the transition..

4

Gill and Nyholm's work (1959), indicates that spectrophotometry

is a better tool for further investigation than magnetic susceptibility

measurements.

Orgel (1955) lists the spectral energy data for several transi-

tion metal ions including cobalt. He begins with Ti(H20)g } which has

only one d electron and only one absorption peak (20,400 cm”*). For his

list of experimental absorption peaks of cobalt(ll) see table 1. The

energy transition of the tetrahalides do not agree with theory presumably

due to hydridization of 3d and 4p levels.

TABLE I

ABSORPTION PEAKS OF COBALT COMPLEXES

Complex Peaks, cm”* - Comments

Co(H20)6++ 8100,19600 Dq = 970 cm”*

Co (NH3)6++ 20200 .

Co(en)3++ 20800,281004 -2 CoCl4 £ 6300,15000 Dti is much less

_2CoBr4 13700 in these

c » v z 12500 tetrahedralCo(CNS)4"2 17000 complexes

An Investigation of a possible CoCl complex was made by Listet

and Rosenblum (1960). They found no change in absorption between zero

and 0.12M chloride. From the shift in electrode potential of the Ag-AgCl

electrode they obtained a formation constant and a AH of 1-0 kcal at

25°C for the reaction, Co++ + Cl" ^ CoCl+ . The ionic strength was

adjusted to 2.0 with NaClO^. It is possible that the absorbance involves

charge transfer from Cl to Co++.“ Smithson and Williams (1958) also

studied ion-association complexes with various anions. These have only

slight affect on the absorbance. Association complexes were found with

chloride, nitrate, and sulfate, but there was no measurable association

with perchlorate. On the basis of the slight change in absorbance, they

developed a slope-intercept equation, and reported a value of 0.37 for

the chloride constant, K^. The ionic strength was adjusted to 7.0 with

LiClO^, and the temperature was 20°. Thiosulfate and thiocyanate formed

more "stable complexes, but these probably contained direct cobalt-ligand

bonds. The thiocyanate had a charge transfer band in the ultraviolet at-2275 run. These authors assumed that CoCl^ is completely formed in 10F

LiCl. Concentrations of CoCl^ and CoCl^" were said to be negligible

because they are unstable in either an octahedral or a tetrahedral

environment.

Ballhausen and Jorgensen (1955) ran reflectance spectra onVsolid CoCl^*2H20 and CoCO^. In both cases the maximum absorption occurred

at 530 nm. The lower wavelength or higher energy indicates an octahedral

environment for cobalt. The same authors assumed the ion in 12F' HC1 to be -2CoCl^ . In concentrated hydrochloric acid they reported absorbance peaks

at 690 nm and 530 nm with molar absorptivities of 600 and 9.5 respectively.

The absorption at 530 nm indicates that some of the cobalt, possibly the

monochloro complex or the simple hydrated ion, is still octahedral.

Cotton and his coworkers have contributed much to the knowledge

of the chemistry of cobalt. Blake and Cotton (1963) made a series of

Born-Haber-type calculations and found the AH for the reaction,—2Co++(g) + Cl“(g) -* CoCl^~ (g), to be approximately -625 kcal per mole at

25°C. The solid compound, Cs2CoCl^, was used as the basis of this cal-- 2culation. It was concluded that the CoCl^ ion should be thermodynam­

ically stable in the solid and gaseous state.

Cotton and the Goodgames (1961, p. 4690), compared the spectra

of [(C^Hg^N^CoCl^, solid and dissolved in methylene chloride, with the

spectra of cobalt chloride in concentrated hydrochloric acid. Since the

spectra were somewhat different, they concluded that the chief ion in

hydrochloric acid solution is tetrahedral 00(^0)01^ . They also found

that cobalt forms a.blue hydroxide complex in 50% sodium hydroxide

solution. __

Holm and Cotton (1959) measured the reflectance spectra and

magnetic moments of solid cobalt halides, such as the following cesium

and quinoline complex salts, Cs^CoCl^ and (CgHgN)^CoCl^. Maximum absorp­

tion of the tetrachloro complex occurred at 700 nm; of the tetrabromo

complex, at 740 nm; and of the tetraiodo complex, at 800 nm. Magnetic

Susceptibility ranged from about 4.7 B.M. for the chlorides to 5.0 for

the iodides. The susceptibility of octahedral cobalt compounds ranged

from 4.8 to 5.3 B.M. (Gill and Nyholm 1959). Note that this overlaps

the tetrahedral range, and in any case shows three unpaired electrons

with spin-orbit coupling.

Katzin (1954) also compared the spectrum of solid Cs^CoCl^ with

that of cobalt chloride in concentrated hydrochloric acid. Both compounds_2have similar spectra; so both were assumed to contain CoCl^ groups.

However, one of the peaks listed for the complex formed by CoCl^ in HC1

is at 533 nm, which is in the octahedral range.

Katzin and Gebert (1950, p. 5464) investigated several of the

cobalt halide complexes in various organic solvents. Some of the absorp­

tion peaks they found are shown in table 2. One can see that there isS

only a slight difference among the absorption spectra of CoCl2 , CoCl^ ,_oCoCl^ . There is an equilibrium set up by the competition of the

solvent, chloride, and water for the ligand positions around the cobalt

ion.

TABLE 2

ABSORPTION PEAKS OF COBALT HALIDE COMPLEXES

Complex \ max, nm Comments

Co (H20)6++ 510 pink

CoX4C12 525,540 X = CH^OH or pyridine

CoX2Cl2 575,615,640,665 X = pyridine, quinoline

575,615,660 X = higher alcohols

575, 630,675 X = acetone, tetrahydrofuran

CoXCl3~ 595,630,665 - X = pyridine, quinoline

595,(625),675 X = alcohols

595,(630),685 X = tetrahydrofuran, acetoneCoCl4“2 615,625,635,665,695 HC1 or LiCl in Acetone

8

Katzin (1952) also made some spectral measurements in the ultra­

violet region. In isopropanol containing a tenfold excess of LiCl, he

assumed that two complexes of cobalt, CoCl* and CoCl^, were formed, but

his figure 3 shows only one type of absorbance curve- at the chloride con­

centrations where these complexes would be expected. Spectra of thio-

cyanates indicated a higher complex, CoCCNS)^ because absorptivity and

the wavelength of maximum absorption increased as lithium thiocyanate wasoadded. In acetone the additional complex, Co(CNS)^ , was obtained. The

4*only thiocyanate complex of nickel in isopropanol appeared to be Ni(CNS) • They (Katzin and Gebert 1953) also compared the reflectance

spectra of cobalt(Il) chlorides in the solid state. Anhydrous crystal­

line CoCl^ is pale blue, and has the CdCl^ structure in which the cobalt

is 6-coordinate. A diagram

(Douglas and McDaniel 1965)

is shown at the right. The

absorption peaks are listed

below.

Salt

CsgCoCl CoCl2-2H20

CoC12H20 CoC1_

X max, nm

675

520

560

600

Cl

Fine (1962) has studied the complexes formed between cobalt(Il)

and chloride ions in acetone solution. His graphs of absorbance at a

fixed wavelength versus the mole ratio of total amount of chloride added

as LiCl to cobalt were essentially linear between integral ratios of 0:1

to 2:1, and 2:1 to 3:1, but nonlinear above a ratio of 3:1. Breaks

occurred at integral ratios of 2:1 and 3:1. An indigo blue color develop­

ed up to a mole ratio of 2:1, and an ultramarine blue developed above

this ratio. These results strongly support his conclusion that CoCl^ and

0 0 0 1 2 " are first formed essentially quantitatively followed by a less_2complete conversion to CoCl^ . Spectra and equilibrium constants were

given for each complex, and are shown in table 3 for the chloride complexes

TABLE 3

ABSORPTION PEAKS OF CHLORIDE COMPLEXES IN ACETONE

X man, nm C max Stability Constants

CoCl2 575 150 P2 > 3 x 109

** 630 (sh.) 225

674 306

CoCl3" 590 237 k 3 > 105

~ 630 150 •

688 455 *

CoCl4"~ 610 (sh.) 248 K4 = 5.4 x 10"2

625 ' 353 .

640 (sh.) 287

667 557

697 612

10

The equilibrium constants were calculated from concentrations,

not activities. The basis for his calculations of the first two constants

is not clear since the data were not included. He also included curvess

for the calculated € versus wavelength in both the visible and very near

infrared region below 2.5 microns.

: Magor and Smith (1968) report that the addition of HgCl^ to an

acetonitrile solution of C o C ^ changes the color from blue to pink. The

authors state that octahedral CoCl(CH^CN)^+ is formed, and that the blue

color reforms upon heating. The di- and trichloro complexes of cobalt

are assumed to be tetrahedral because the spectra are similar to that of

the tetrachloride. The absorbance peaks and absorptivities of the three

complexes in acetonitrite are similar to those in acetone listed by Fine

(1962).

Fine (1965) also has investigated the halide complexes of

nickel in acetone. He found that the tetrabromide complex of nickel canibe prepared in acetone, provided excess LiBr is added. The stability

- • ‘ - -2constant for the final step, NiBr^ + Br ** NiBr^ , as determined by the-2slope-intercept method is 1.0 x 10 . The only bromide complex in

aqueous solution is octahedral Ni(H2 0 ),-Br+.

Nickel, like cobalt, forms tetrahedral chloride complexes, such

as (R^lOgNiCl^, in the solid state (Cotton, Goodgame, and Goodgame 1961,

p. 4161). In polar solvents, even nitromethane, these complexes disso­

ciate. Magnetic measurements show two unpaired electron's.

Peter Pauling (1966) prepared the tetrahedral nickel complex, *

[(CgHs)^CH^AsJ2^ iC1^, which is stable and soluble in nonpolar organic

solvents. Similar compounds were made from divalent manganese, iron,

11

cobalt, copper, and zinc. All are isomorphous with the nickel complex

except the copper compound. Thus it is shown that the first row of

divalent transition metals all can form tetrachloro complexes. In theUsolid state the copper complex can exist in the square-planar as well as

the tetrahedral form (Willett and Liles 1967).

Cotton and his coworkers also made spectrophotometrie measure­

ments on the solid tetrahedral complexes of cobalt pseudohalides, such as

the thiocyanates (Cotton et al. 1961, p. 4157); cyanates and azides

(Cotton and M. Goodgame 1961, p. 1777); and selenocyanates (Cotton et al.

1962, p. 565). The absorption spectra and magnetic moments are similar

to those of the tetrahalides.

Very few references mention any use of electrical measurements

for determining the nature of cobalt chloride complexes. However, Miss

Wormser (1948), by using a modified Job’s method (1928), showed that in

acetone there was maximum conductivity when the ratio of LiCl to CoCl^ was

1:1. This suggested the presence of the complex, Li^CoCl^ . In hydrochor--i I ,|, wic acid the ions, Co , CoCl , and CoCl^ , but not CoCl,,, were assumed to

be present. The blue color was assumed to be due to the anion, CoCl^".

During electrolysis of HC1 solution, cobalt moved toward both the anode

and cathode, but much more toward the cathode.

Other inconclusive electrical measurements were performed by

Moore and Kraus (1952). First they found that the maximum absorption of

cobalt on the anion exchange resin, Dowex 1, occurred from 9F HC1 solution.

Here the concentration of CoCl^” was thought to be at its highest. In

electrophoresis experiments, cobalt(ll) began to migrate to the anode at

concentrations of HC1 above 8F, but no data were given.

12

Reports on solvent extraction studies of cobalt chloride

complexes were also scarce. One was made by Lindenbaum and Boyd (1963),

who used tri isooctyl amine to extract the chlorides of Mn(Il),- Fe(lll),

Co(ll), Ni(ll), and Cu(ll). The spectra of the organic extracts of

ferric, cobalt, and copper chlorides were practically the same’ as the;

spectra in concentrated hydrochloric acid. Therefore, these chlorides in

hydrochloric acid were assumed to be of the MCl^ type. The only chloride

complex of nickel in aqueous solution appeared to be NiCl+. Rutner (1961)

performed similar studies with Co(ll) and Fe(lll) chlorides, and reached

the same conclusions. The difference was that the chlorides were absorbed

on solid amine-type resins rather than by extraction with amine solutions.Good and Srivastava (1965) carried out another solvent extrac­

tion study. They used amines of the type, R^NCl or R^NHCl, to extract

cobalt(ll) from 8.5F LiCl or HC1. Since the ratio of amine to cobalt-2was about 2:1, the cobalt complex was assumed to be CoCl^

Sato (1967A) used tri-n-octylamine dissolved in benzene to

extract cobalt chloride from hydrochloric acid or lithium chloride

solutions. The latter was preferred, where the partition coefficient

into the organic solvent is about 30. The maximum extraction of cobalt

takes place from 9 to 10 M chloride solutions. The proposed reaction is

CoCl2(aq) + 2R3NHCl(org) ** (R3NH)2CoCl4(org).

Some workers tried ion exchange methods in their effort to

identify the cobalt chloride complex. One group (Kraus et al. 19.55) used

the anion exchange resin, Dowex 1, in their study of the absorption of

metal chlorides. They found that the chlorides of Sc(lll), Fe(lll), Co(ll),

13

Zn(ll), Ga(lll), Au(lll), and Be(ll) are absorbed more strongly from LiCl

than from HG1 solutions.

The same- anion resin was used by Herber and Irvine (1956) ino

their study of the formation of nickel chloride complexes. In 12F HC1,

nickel chloride ran through the column almost as fast as sodium chloride.

From absorption data the plot of log 7.T against log Y± for HC1 at 25°C

gave a slope of -1.0. This fact indicates that NiCl+ is the principal

complex.

The same authors (1958) continued work on the absorption of

cobalt chloride from hydrochloric acid solution. They suggest without

proof that even though cobalt is absorbed on an anion resin, it need not

exist in the aqueous phase as a complex anion, but that neutral CoCl^

reacts with the resin to form an anionic complex. The plot of log %T

against Y^ is linear up to a concentration of 9F for HC1, and the slope

is -1.95. This fact suggests that there are two chlorine atoms percobalt, and the complex can be either CoCl^'AH^O or CoCl^^^O. They

_2give an equilibrium constant of 5.3 x 10 for the reaction in 9F HC1:

Co'H '(aq) + 2C1" ^ CoCl2(aq).Herber and Irvine also list activity data for hydrochloric acid in

concentrations from 6 to 11F.

Coleman (1966) placed CoCl^-HCl solutions on both cation and

anion exchange resins. Samples low in chloride, when dried on an anion

exchanger and then exposed to gaseous HC1, also produced a blue color.As a closing thought the author hints that there just may be no anionic

complex of cobalt in concentrated hydrochloric acid.

14

Morris and coworkers (1965) list the association constants,

shown in table 4, for several divalent transition metal chlorides.

TABLE 4

ASSOCIATION CONSTANTS OF TRANSITION METAL CHLORIDES

Ion Log 0 ^ Log 02 Log 03 Log 0^

Mn 0.59 0.26 -0.36 -

Co 0.69 0.51 - -Ni 0.23 -0.04 - -

. - Cu 0.98 0.69 .0.55 0.0Zn 0.73 0.49 -0.19 +0.17

The absence of data.for cobalt beyond P2 1® interesting. The data also

confirm that all chloride complexes of nickel beyond the monochloride are very weak.

The literature contains many references to the three-way

competition of water, organic solvent, and halide ions to be ligands of

the transition metal ions. For instance, Buffagni and Dunn (1960) state_2that nearly all oxygen-containing organic solvents convert NiCl^ to

Ni(solvent)g++. In the case of cobalt there are often intermediates. Both

octahedral and tetrahedral complexes, which are partially chlorinated, may

exist in equilibrium. A solvent used by the authors was dimethylformamide,

and it was reported to form chiefly Co^ffOCl^”.

15

Pasternack and Plane (1965) studied the competition of water,

acetone, and ethanol for coordination sites around the cobalt(ll) ion.u

(See table 5.) Noncomplexing perchlorate was used as the anion. From a

purely statistical viewpoint, there are six positions where the first

solvent molecule may enter, and one position from which it may leave.

Therefore, the statistical contribution to the equilibrium constant, K^,

would be 6. The actual values for acetone and ethanol are around 3,

or half of the value expected from statistics alone, with the rate

constants in the forward and reverse directions being equal. For the

second sovent molecule to coordinate to the cobalt ion, there are five

positions wher-e- it may enter, and two positions from which it may leave,

corresponding to a statistical of 2.5. The actual Kg was much less

than this, and this fact was attributed to the distortion of the octa­

hedral- structure of the complex. *

TABLE 5

ABSORPTION DATA FOR SOLVENT COMPLEXES OF COBALT(II)4

Solvent X max, nm A max *L K2

Water 513 A.90 -» -Acetone 515 10.51 3.6 0.015

Ethanol 517 8.30 2.9 0.006

* 16

Matwiyoff (1966) found that the optical spectrum of cobalt

perchlorate is practically the same in any polar solvent. He used water,

methanol, and dimethylformamide. Coordination is always sixfold and

through oxygen.Scaife and Wood (1967) studied equilibria of the type,

tsolvent)^ + X" ^ WXg(solvent)" + solvent, (X = Cl or Br; M = Co or

Ni). The solvents used were water and alcohols. The tetrahedral species

appeared to be favored at around 85°C, while the octahedral species was

favored at room temperature. The authors suggested the presence of

CoCl^HgO)*" and CoCl^CH^OH)“ In 12F lithium chloride at room temperature.

In n-butanol, nickel forms only NiBrg^^H^OH) or ^ in excess LiBr.

Several investigators have written about mixed complexes of

cobalt(ll) with halide and either pyridine or quinoline. Katzin (1961)

measured the energy change when one such complex .goes from the octahedral

to the tetrahedral configuration. He used the reaction, '

CoCl^py)^ 54 C o C l g ^ y ^ + 2py; where py = pyridine.

At 38°C the AH = + 13.4 kcal/mole, and AS = •}• 36.7 e.u. The equilibrium oconstant for the reaction as written is 0.04 at 38 C. - A higher tempera­

ture drives the endothermic reaction to the right. In methanol the

following reaction was said to take place at 45 to 58°C.

2CoClo(CH-0H). ** CoCl(CH„0H)c+ + CoCl-CH.OH" + 2CH.0H.2 3 4 3 5 3 3 3The absorbance peak of the trichloride complex occurred at 592 nm.

King, Koros, and Nelson (1963, 1964) studied the effect of steric

hindrance upon the type of complex formed. Quinoline and 2-methyl

pyridine formed only tetrahedral complexes of the type C 0Q 2CI2 with

cobalt(ll). Pyridine and isoquinoline, which are not sterically hindered,

17

can form either CoCpy^0^ or Co^Py^4C12* The latter is octahedral. Increasing the size of the halide ion also favors the.stability of the

tetrahedral complex.The Goodgames (1963, p. 207) prepared mixed halide-quinoline

complexes of nickel and cobalt with the general formula, The

cobalt(ll) complexes tested are all tetrahedral. The blue chloride and

bromide complexes of nickel are assumed to be tetrahedral because they

are paramagnetic. There is also an insoluble yellow chloride complex

which has become a polymer due to chloride bridging. It appears to be

octahedral. The iodide appears to be square-planar because it is

diamagnetic.

Clark and Williams (1965) also worked with pyridine complexes.

They prepared both tetrahedral and octahedral pyridine-chloride complexes

of cobalt(ll).

Bertrand and Plymale (1964) made an interesting study of a

pyridine N-oxide complex. The empirical formula is CoL^Cl^, where

L = C^Ht-NO. This was first thought to be (CoL^Cl)+Cl . However, spectra

$nd magnetic measurements (4.75 B.M.) indicated a tetrahedral structure

with all of the chloride (or bromide) bound to cobalt. Infrared measure­

ments suggested octahedrally bound, organic ligands. They concluded that

the compound is (CoL^)(CoCl^). An analogous cadmium bromide, (CoL^)(CdBr^)

was prepared, and this helped to confirm that the anion was a tetrahalide complex.

Cotton, Faut, and Mague (1964) prepared some mixed halide-

thiourea complexes, such as Co^SNgH^^Clg and Co^SNgH^gBr^, in which

the coordination is through sulfur. Both are tetrahedral and are

nonconductors when dissolved In nitromethane.Apparently, the only known fluoride complexes of cobalt are

found in the solid state. In the only literature reference noted

(Crocket and Grossman 1964), there is mentioned the formation of

where M may be divalent Mn, Co, Ni, Zn, or Cd.

Some work In fused chloride systems was done by Gruen and his

coworkers (0ye and Gruen 1964, 0ye and Gruen 1965, Angell and Gruen 1967).

In fused potassium chloride both nickel and cobalt are reported to be

four-coordinate or tetrahedral. In fused aluminum chloride nickel and■f*3cobalt are six-coordinate or octahedral. The small A1 Ion takes some of

the negative ctfarge away from the chlorides, so that the chlorine atoms

do not repel each other so much, and more of them can gather around the

transition metal cation. In fused zinc chloride in which the polarity is

intermediate, cobalt ion is tetrahedral, while nickel is octahedral in

that solvent salt. When equimolar amounts of KC1 and AlCl^ are mixed, a

eutectic, KAICI^ is formed. Cobalt chloride precipitates from this

eutectic. If more KC1 is added, tetrahedral I^CoCl^ is formed.

Other Cobalt Complexes

Up to this point only tetrahedral and octahedral complexes of •

cobalt(ll) have been discussed. In general, cobalt forms six bonds with

oxygen or nitrogen-containing ligands. The larger halogen atoms usually

permit room for only four bonds. The negative charge on the halide ions

also tends to push them apart. Under certain conditions cobalt can also

form five-and eight-coordinate complexes.

19

Four-coordinate square-planar complexes of cobalt(ll) are

unusual, but they do form when the geometry of the ligand demands it,

as in porphyrins. Even here weaker bonds are formed in the trans posi-*

tions with anions or solvent. One example of a square-planar complex is

with o-phenylenebisdimethylarsine (Rodley and Smith 1967A, Einstein and

Rodley 1967). Anions including perchlorate and nitrate occupy the trans

positions to form an octahedral complex in the crystalline state. The

anions break away and ionize in a polar solvent.

A few keto-amines can also form square-planar complexes

(Everett and Holme 1965). Magnetic measurements show only one unpaired2electron, which indicates dsp bonding. In the accompanying figure, R

can be or^H^.

CIO, CH_/ ^0 . 'sc— o

/Ask 8 v HC^ ^ Co/2iCo ^ C N/ \A s^ I "As' R H

'cio3

In an oxygen environment cobalt(ll) can be tetrahedral if the_2ligand is small per unit charge as in Co(OH)^ (Cotton, Goodgame, and

Goodgame 1961, p. 4690). Another case occurs when the ligands are large

and bulky. One example is the diacetylacetonate which is tetrahedral

when dissolved in a nonpolar solvent (Cotton and Elder 1966, p. 423).

The anhydrous solid, however, is a tetramer, and by forming oxygen

bridges, the cobalt becomes octahedral (Cotton and Elder 1965, p. 1145).

20

Another tetrahedral complex with cobalt-oxygen bonds is

[(CgHij^As^CoCOgCCF^)^ (-Bergman and Cotton 1960, p. 1420). Here the

ligand is only monodentate, but the bulky trifluoracetate ligands and

large cations help to stabilize the tetrahedral structure.

One of the few examples of a five-coordinated cobalt(ll)

complex is described by Bertrand and Plymale (1966, p. 879). Dibromo-

trls(diphenylphosphine)cobalt(ll) is described as having a trigonal

bypyramid structure. Both bromides and one phosphorus are in the

equatorial plane. Magnetic measurements show one unpaired electron,3which indicates dsp bonding.

Some of the most interesting complexes of cobalt are the

nitrates. Cotton and his coworkers have investigated many of them, and

find that the nitrate ion is often bidentate. One example is a nitrate-

phosphine oxide complex, Co[(CH3)3P0]2(N03)2 (Cotton, D. M. L. Goodgame,

and Soderburg 1963, p. 1162). This compound is nonconducting and soluble

in nonpolar solvents such as chloroform. X-ray studies show six-oxygen

atoms around each cobalt. The maximum absorption occurs around 550 nm.

This is a second indication that the compound is octahedral, and that

each nitrate group supplies two oxygen atoms for bonds with cobalt.

Other nitrate complexes investigated were [(CH^^N^CodtfO^)^

(Cotton and Dunne 1962, p. 2013) and [(CgH3)^As]2Co(N03)^ (Bergman and

Cotton 1966, p. 1208). Both compounds are soluble in nitromethane. The

former has a maximum absorption at 530 nm, and may be octahedral. The

latter was shown to contain cobalt with a coordination number of eight

so each nitrate group must be bidentate.

21

The investigators, Addison and Sutton (1964), discovered the

complexes, CoCNO-^N^O^ and CoCNO^^CN^O^)^. The former structurally is

N0 +Co(N0 2 )3 ~» which is octahedral with bidentate nitrate.ligands. The

compound decomposes at about 105°C to anhydrous cobalt nitrate, which is

ionic.The formation of nitrate association complexes was investigated

by Katzin and Gebert (1950, pp. 5451 and 5455). Part of the water of

hydration could be replaced by an organic solvent, but the total,

coordination number remained at six. In anhydrous acetone or alcohol

the final compound seemed to be neutral CoOTO^^C s o l v e n t ) T h e addition

of tetrabutylammonium nitrate produced evidence of a Co^Og)^ complex (Katzin and Gebert 1950, p. 5455). Their method used spectral data,

which was analyzed by methods developed by Job (1928) and Vosburgh and

Cooper (1941). Observe that only polar solvents were used. Biagetti and

Haendler (1966, p. 383) prepared complex pyridine nitrates of cobalt(ll).

The most stable was Co(py)^(N0 ^ ) 2 which was octahedral and contained one

mono and one bidentate nitrate ligand. The compound was soluble in

chloroform. Adding more pyridine formed C0 (py)^(N0 ^ ) 2 and Co(py)g(N0 2 )2 «

Even a few cobalt perchlorate complexes have been' made. One is

the diarsine derivative already mentioned (Rodley and Smith 1967A,

Einstein and Rodley 1967). Another was investigated more thoroughly by

Cotton and Weaver (1965). The compound is C o ^ H ^ S ^ ^ ) 2801^ ) 2 (0 1 0 ^) 2 and is shown at the right. The Co-S bond CH_ C }°3 9H 3

|3 I Ilength is 2.29 A. The Co-0 bond length 3 ?^2

1 ^ Co 1is 2.34 A, which indicates a weak, but h C ~ q 1 " ^ S __ CH

2 | 0 1 2definite bond. CH^ ^ 3

22

Cobalt can also form complexes in the +1 and zero oxidation

states. These are generally with TT-bonding ligands such as carbon monox­

ide and isonitriles. One example' is pentakis(methylisonitrile)cobalt(l)

(Cotton, Dunne, and Wood 1965, p. 318), which has the trigonal bipyramid

structure. The Co-C bonds have a bond order of about 1.5.

Halide Complexes of Certain Platinum Metals

Although cobalt is more like its horizontal neighbors in the

periodic chart, it was thought that a literature study of the halide

complexes of the vertical neighbors might throw light upon the composi­

tion of cobalt halides. The elements below cobalt, rhodium and iridium,

are somewhat different in that the most stable oxidation state is +3.

Since the ions are larger, higher coordination numbers may be expected.

The most complete report on the chloride complexes of rhodium(lll)

was written by Wolsey, Reynolds, and Kleinberg (1963, p. 463). They+3 —3prepared all of the mononuclear complexes from Rl^^O)^ to RhCl^” • The

yellow hydrated rhodium(lll) ion may be prepared by dissolving Rh(OH)^ in

dilute perchloric acid. The addition of each chloride causes a slight

increase.in the wavelength of the absorption band. The hexachloride

anion is dark red. The authors (Wolsey, Reynolds, and Kleinberg 1963)

prepared the various chloride complexes by boiling the perchlorate

solution with hydrochloric acid of the required concentration, which is

given in their paper. Rhodium(lll) is like chromium(lll) and cobalt(lll)

in that it is changed very slowly at room temperature from one complex to

another. Kristjanson and Lederer (1959, p. 245) reported the presence of_5

polynuclear complexes, such as Rh^Cl^ in 6F hydrochloric acid.

Rhodium also forms complexes with sulfate (Shukla and Lederer

1959, p. 255) and oxalate (Shukla 1959, p. 333). The oxalate complex is

a yellow anion, which forms the salt, K^Rh(C^O^)^. In sulfate solution

rhodium forms both yellow cationic and red anionic complexes, the latter

being formed only by boiling with concentrated sulfuric acid. Rhodium(lll)

forms alums similar to those of chromium(lll).

Rhodium(l) and iridium(l) form several square-planar complexes

with TT-bonding ligands such as carbon monoxide. Examples (Vallarino 1965)

are the dimer, [Rl^Co^Cl],, wi-bh chloride bridges, and the anion,

RhCCO^Clg”. The chloride can be substituted with bridges of oxygen atoms

contributed from acetate or nitrate (Lawson and Wilkinson 1965, p. 1900).

The bridges can be broken with triaryl phosphines and certain amines.

Iridium(l) can even form a complex with molecular nitrogen, IrL^Cl^, where

L is triphenyl phosphine (Coilman 1968).

The most thorough quantitative work on iridium(lll) chloride

complexes has been done by Garner and coworkers (Paulsen and Garner 1962,

p. 2032; Chang and Garner 1965, p. 209; El-Awady, Bounsall, and Garner

1967, p. 79). They first (Paulsen and Garner 1962) started with the-3 -2reaction, IrCl^ + 1^0 ** Ir(H20)Cl^ + Cl , which was slow even in

boiling water. The second and third papers dealt with the replacement of

the second and third chlorides, respectively, with water. The final

product was neutral I^H^O^Cl^- Liberated chloride ion was easily

determined by titration with silver nitrate. Kinetics of these reactions

were discussed in some detail. The presence of nitrate seemed to hasten

the replacement of chloride, so perhaps an intermediate nitrate complex

was formed. As each cloride was replaced by water, there was a shift in

• . . . 24*

absorption toward shorter wavelengths. Oxidation to Ir(lV) caused the

complexes to have absorption of greater magnitude and at higher wave­

lengths.Another indication of the high stability of iridium chloride

complexes is that they are not reduced completely to the metal by zinc

or magnesium even after the solution is boiled for one hour (Beamish 1966

p. 57). Rhodium(lll) is reduced under these conditions.

MacNevin and his coworkers (MacNevin and Crummett 1954,' p. 323;

MacNevin and McKay 1957, p. 1220; MacNevin and Dunton 1957, p. 1806)

separated some of the complexes of the platinum metals by means of ion

exchange. In neutral solution rhodium forms an insoluble hydroxide and

does not move on an ion exchange resin. Platinum forms an anionic-2hydroxy complex, Pt(OH)^ . The addition of EDTA forms an anionic

-2 -3complex, probably PdY or PdYOH - , with palladium. The very stable-3iridium chloride complex, IrCl^ , apparently remains unchanged. During

electrophoresis (1957, p. 1806) in neutral solution containing EDTA, the

rate of migration toward the anode is of the order Ir > Pt > Pd > Rh.

In dilute acid, pH 2.8, only rhodium forms a stable cation in chloride

solution. MacNevin and his coworkers (1954, p. 323; 1957, p. 1806) made

the qualitative assumption that the chlorides of irldium(lll) are less

reactive kinetically than those of rhodium(lll).

Lederer (1958), p. 279) and Shukla (1958, p. 457) also used

electrophoresis and ion exchange in their work. Lederer showed that

upon aging in dilute HC1 or HBr, a mixture of cation and anion complexes

of rhodium were formed. In nitric acid there were two complex cations; .

the chief *one was Rh(Ho0)_NO_++ with some Rh(H„0),+^. Rhodium(lll)Z o o Z o

. . ' 25

cations were said to form irreversible complexes with the sulfonic acid

groups of Dowex 50 resin, but no data were given. Electrophosesis (Shukla

1958, p. 457) of rhodium(lll) in dilute perchloric acid resulted in three

bands. The largest was the hydrated ion, RhCl^O)^**^. The others were the

mono and dihydroxy complexes. Apparently no perchlorate complex was

formed.

Careful experiments on palladium chlorides were performed by

Mrs. Weed (1964) in these laboratories. By means of spectrophotometry,

she determined the four individual constants as four chlorides were

successively added to palladium(ll).. The first three constants were

determined by using Bjerrum's (1944) method of corresponding solutions.

The fourth stability constant, K^, was determined by one of the modifi­

cations of the slope-intercept method (Whiteker and Davidson 1953).

Palladium(ll) is different from cobalt(ll) in that there is only a

gradual change in color as each chloride is added. The chief reason is

that the structure remains the same, namely square-planar.

Summary/

It appears that cobalt(ll) can form an amazing variety of

complexes. Although tetrahedral and octahedral are the most common, there

are examples of square-planar, trigonal bypyramid, and even eight-

coordinate complexes. Cobalt(ll) usually bonds through oxygen, but there

are many ligands which bond to cobalt through nitrogen, carbon, sulfur, phosphorus, and halogens.

The only agreement in the literature about the nature of the

chloride complexes in concentrated hydrochloric acid seems to be that they

26

are blue and absorb in the 650-700 nm range. Various complexes listed

are Co(H20)3C1+ , Co(H20)5C1+ , CoCl2(aq), CoCl+CoCl3" , Co(H20)C13", and _ 2CoCl, . Most investigators agree that the blue color is fully develop-

ed in 9-10F HC1 or LiCl. Then some add that there is also absorption at

525-530 nm, a fact which indicates that an octahedral complex is still

present. Bromide and iodide complexes absorb in the 700-800 nm range.

Nitrates, perchlorates, sulfates, and dilute solution of chlorides

absorb in the 515-530 nm range. The complex is said to be the hydrated

ion Co(H20)g . Oxygen-containing organic solvents may replace water as

a ligand.

CHAPTER II

ELECTRICAL AND ION EXCHANGE EXPERIMENTS

sIn view of the fact that there are so many conflicting

interpretations about the structure of the blue cobalt chloride complex,

it was thought that electrical migration and related experiments could

determine if the complex is positive, negative, or neutral. Then it

should be easier to interpret the complicated changes in absorption

spectra.During electrical migration an ion may move to a new environ­

ment. In this new environment complications often arise because the ion

or complex is equilibrating to the new conditions. The^e new equilibrium

conditions can form a new complex with a different charge, and thus

change the rate of migration. For instance a spot of cobalt chloride

moving in electrophoresis is moving into an environment containing no

Cobalt ion. If the cobalt chloride complex were binuclear, the complex

would tend to dissociate where no cobalt is present. The following is a-2 -hypothetical reaction: Co^l^ “* 2CoC1'2 . The singly charged mononu­

clear complex would move only half as fast as the binuclear complex. The

following is a possibility if the chloride concentration is reduced,

possibly by evaporation of hydrochloric acid: CoCl^” **♦ C o C ^ + Cl“. The

neutral CoCl^ would not move at all under an electrical potential. It is

the movement at the beginning of the experiment that shows the charge of

the original complex.27

28

Electrical Migration Experiment

In this experiment a solution of CoCl^ in 12F LiCl was

dispersed in an agar gel at the bottom of a glass U-tube as shown in

figure 1. The side arms were filled with 12F LiCl, and platinum wires

connected to the terminals of a lead storage battery were immersed in

the solutions. After two hours of electrolysis the absorbance in the

anode compartment at 700 nm was 0.014. In the cathode compartment it

was nearly six times as great or 0.080. This observation indicates++ 4*the presence of some cationic complex, probably Co or CoCl , but

very little anionic complex. When there was no current, there was no

detectable diffusion of cobalt ion from the agar gel into the lithium

chloride solution.

Electrophoresis

A second study of electrical migration was that of electro­

phoresis on paper. Two dishes containing electrolyte were place at

the ends of a water-filled tank 29 cm long and kept at 25°C. A

platinum wire was placed in each solution and each wire was connected

to one of the terminals from a high-voltage direct-current power source.

Two spots, one of cobalt and one of copper chloride, were placed on the

center of a strip of No. 1 Whatman paper. Then the paper was wet with

the electrolyte. The ends were dipped into the two solutions, and this

completed the electrical circuit. A cover was then placed over the

whole apparatus except the power source. After two hours of electro­

phoresis at a current of 50 milliamperes, the position of each colored

29

. FIGURE 1o

ELECTRICAL MIGRATION APPARATUS

CoCl„ in LiCl

LiCl

. - ■ . . . 30

spot was measured. The summary of the measurements appears in table 6.

A negative sign indicates movement toward the cathode, the negative

■ electrode.

TABLE 6

MIGRATION UNDER ELECTROPHORESIS

Chloride MolarityMigration in Centimeters in Two Hours

Co in HC1 Co in LiCl Cu in HC1 Cu in LiCl

1 - 1.9 - - 1 . 8 -

3 - 1 . 2 - - 1 . 2 -

6 - 1 . 0 - 1.3 - 0.4 - 0 . 8

7 - 1 . 0 - 0.9 - 0.3 - 0 . 2

8 - 0 . 8 ‘ - 0 . 6 - 0.3 + 0 . 1

9 - 0.7 0 - 0 . 2 + 0.3

1 0 - 0.5 + 0 . 1 0 . + 0.7

11 - 0.4 + 0.5 0 + 0.7

1 2 - + 1 . 0 .+ 1 . 0

In hydrochloric acid up to 11F the cobalt always- migrated toward

the cathode. The small amount of migration, however, in 10 or 11F HC1

indicates that most of the cobalt is probably bound in a neutral complex.

Practically all the copper appears to be in a neutral complex because

there is no migration at all. Concentrated or 12F hydrochloric acid

could not be used because the paper disintegrated. However, before the

31

disintegration both cobalt and copper started to migrate toward the

anode. Thus, some anionic complex appears to exist in concentrated

hydrochloric acid.In lithium chloride above 10F both cobalt and copper(II)

complexes migrate toward the anode. It appears that some anionic

complex is present in LiCl but not in HC1 at these concentrations.

For cobalt the absorption in the 650 to 700 nm range (described in

detail in chapter three) is somewhat greater in HC1 than in LiCl of the

same concentration. For copper there is a slight difference in the

spectrum around 380 nm, so two different complexes may be present. The

neutral complex, where there is little or no movement, is probably

MClg^gO^- The cobalt complex is presumed to be tetrahedral, but the

copper complex may be square-planar (Eswien et al. 1967A).

The movement toward the anode in 11 and 12F lithium chloride

solutions is probably due to the trichloro complex, MCl^I^O)”. In no

case did any metal spot divide into two. The movement appeared to

represent an average movement of all complexes present.

During the electrophosesis experiments it was noted that the

migration distance is not a linear function of*time. About 75% of the

migration took place during the first hour. After the current was

shut off at the end of the second hour, the spot often moved two or

three millimeters in the reverse direction, but the reason was not

investigated. However, no spot ever reversed its direction as long

as the current was on. The final measurement was .recorded before the

power was turned off.

It should be mentioned that because of evaporation, the

chloride concentration in the paper may not have been the same as in

the electrolyte. Volatile hydrogen chloride leaves 10F solutions before

water does, so the actual acid concentration in the paper was probably

less than 10F. In the case of lithium chloride, the alkali salt does

not evaporate, but the water does. Therefore, the salt concentration

in the paper may have been more than the nominal amount. This may

explain the discrepancy of the difference in migration in the two

electrolytes. The differences in hydration and the activity of the

hydrogen and lithium ions may also have been factors.Because of these complications, the data cannot be considered

quantitative. Their chief significance is to determine the sign of

the electrical charge of the complex that migrates.

In summary it appears that in 10F chloride solution copper

is present mostly as a neutral dichloro complex. Cobalt appears to be

present mostly as a neutral complex, but a cationic complex is also

present in small amounts. When the chloride concentration is less

'than 10F, copper and cobalt exist partially as cationic complexes.

Electrophoresis alone does not tell whether these are hydrated M++ or

MCI . When the chloride concentration is more' than 10F, a little

anionic complex appears to be present.

Ion Exchange Experiments

Ion exchange is another tool which may show whether a complex

is positive or negative. Both anion (Bio-Rad Ag 1-X8) and cation

(Dowex 50W-X8) exchange resins were tried. In general the flow of a

33*

cationic complex is retarded by a cation exchange resin, while an anion

is- retarded by an anion exchange resin. Anions are attracted to,the

quaternary ammonium cations, which are fixed in the resin by organic

chemical bonds. The fixed group in a cation exchange resin is usually

a sulfonic acid group, SO^H, which is bound to the organic skeleton of

the resin. The hydrogen may ionize as a solvated proton and thus be

exchanged for a metal cation. . As in electrical migration, one must

remember that an ion is moving to a new environment and the new equili­

bria established may change the rate of flow through the resin. A

neutral group usually flows through either resin unchanged, but it may

be retarded if it is polar.

In the ion exchange experiments a few drops of copper or

cobalt chloride were placed on the column, which had already been washed

with 10F hydrochloric acid. The colored metal complex could be followed

as a band through the column. With cobalt(ll) on a cation exchange

column, part of the green complex remained right on top of the column.

Some of the green color moved several centimeters down the column as 10F /HC1 was added. Thus some of the cobalt appears to be ‘present as a

cationic complex, but there is an equilibrium which produces neutral or

even anionic complexes. On an anion exchange resin the green band moved

slowly down the column, and there was no division. There apparently is

no strong anionic complex in 10F HC1, but there may be a weak one.

For the yellow copper chloride complex on a cation exchange

resin, the color moved the equivalent of one milliliter for every 20 ml

of 10F HC1 that was added. This could indicate a neutral complex or a

. 34

weak cationic one. In the anion exchange column the yellow band hardly

moved at all, and this fact suggests an anionic complex.

Since most ion exchange work is still empirical, only generalo

conclusions can be drawn from these experiments. However, in 10F

hydrochloric acid, cobalt(ll) appears to show an equilibrium between a

cationic complex and a neutral one. The presence of an unstable anionic

complex cannot be ruled out. In the case of copperCII) there are hints

of both cationic and anionic properties, so probably the principal

species is neutral. These facts agree with the electrophoresis

experiments. The probable equilibria can be expressed as follows with

the water of hydration omitted:

Co++ + 2Cl"" ** CoCl2

or,

CoCl+ Cl" 5s CoCl2

and

CuCl+ + 2C1" ^ CuCl2 + Cl" 5* CuCl3".

CHAPTER III*

SPECTROPHOTOMETRY OF COBALT(ll) AND COPPER(II)IN AQUEOUS CHLORIDE SOLUTIONS

Introduction

For studying inorganic compounds and complexes, one generally

uses the visible and ultraviolet portions of the spectrum. Here the

energy transitions are electronic. Energy is absorbed as an electron is

raised to a higher orbital, such as from 3d to 4p, or from a bonding to

an antibonding orbital. Each transition occurs at a specific wave­

length, but the absorption peaks are rounded due to variable interatomic

forces and because vibrational and, in the case of gases, rotational

transitions are superimposed on the electronic transitions.

With transition metal ions electronic transitions are of two

general types. In the first, the absorption is relatively weak, ‘and

the color is rather pale. The transition occurs between two d levels

which have been split by adjacent ions or polar molecules. The absorp­

tion is weak because the energy transition between two d levels is

"forbidden." A second type of electronic energy transition, known as

"charge transfer," occurs where an electron jumps from one atom to

another. This is permitted in accordance with electronic energy

selection rules, and it often causes strong absorption and deep colors.

35

- ' ' 36

This type of absorption occurs in compounds of a transition metal with a

halogen, in which the bond is partly ionic and partly covalent.

In octahedral, complexes the degenerate d levels are split into

two energy levels. Two of the orbitals, indicated by the symbol, e ,Spoint right toward the ligands, and are of higher energy. The three

remaining orbitals, fc2g» P°int between the ligands, and are of compara­

tively low energy. The observed color of most transition metal ions is

due to the absorption of energy when an electron jumps from a tg^ to an

e orbital.8

With tetrahedral bonding all the d orbitals point between the

ligands, and the separation in energy between e^ and t2 ^ orbitals is then

not nearly so great. Therefore, less energetic radiation, or a higher

wavelength, is required to make an electron jump to a higher d orbital.

Cobalt(ll), which forms both types of complexes, shows this

energy relationship very well. Complexes in which the bonding occurs

through oxygen atoms are nearly all octahedral, and absorb in the 515 to

535 nm range. Complexes with the larger halogen atoms are often tetra-

jiedral, and the absorption is in the less energetic 650-700 nm range. ■* . •

The absorbance bands of the halide complexes in the 650-700 nm range are

very large, possibly because of charge transfer phenomena which are

permitted by electronic energy selection rules.

In general, when, a shift occurs in the wavelength of the

absorbance peak or in the molar absorptivity, a new complex has been

formed. A slight shift suggests that the outer environment with the

solvent has changed, but that the metal-ligand bonds themselves have not.

37

Two instruments, the Cary 14 and the Beckman DB, were used for

obtaining absorbance data. The Cary 14, when available, was used to plot

the entire visible, and sometimes- the ultraviolet spectrum. When the

Beckman DB was used, the absorbance was recorded at intervals of about

25 nm. When a maximum or minimum was observed, this point was also

recorded. Afterwards, the points were graphed. The zero absorbance was

adjusted with blanks at one end of the significant spectrum. The refer­

ence cell was filled with pure solvent or acid. Usually a one-centimeter

cell was used, except when dilute solutions or those of low absorbance

required a ten-centimeter cell. The temperature was 25 dh 1°C.

Spectrophotometry With Common Anions

Dilute solutions of cobalt(ll) salts with all common inorganic

anions have the same absorbance curves. The absprbance maximum is at

515 nm and the molar absorptivity is 4.9. The hydrated ion, CoCl^COg**

is assumed to be present. The pink color is visual evidence that the

cobalt is octahedral, because it takes more energetic light (green) to

force a t^ electron to the higher e^ energy level.

In the presence of inorganic acids, other than hydrochloric,

there is only a slight change in the absorbance curve. Curves in the

presence of nitric and sulfuric acids are shown in figures 2 and 3. With

nitric, perchloric, and sulfuric acids, there is practically no change in

absorbance up to a concentration of about 3M. When these acids become

about 4M, the absorbance begins to increase slightly and the effect of all

three acids on the absorbance of cobalt(ll) is the same until their concen*

tretion reaches 9M. At this point the water content of perchloric and

38

FIGURE 2

ABSORBANCE OF Co(ll) IN HN03

(C . ss 0.010F;o HNO^ molarity as indicated)

1.00

0.80

0.60

0.40

0.20

600 450 nm550 500

39

FIGURE 3 ABSORBANCE OF Co(ll) IN I^SO^

(CCo =5 0.010F; H^SO^ molarity as indicated)

OJ

00

OJOJ

OO O o o

650

600

550

500

450

nm

sulfuric acids becomes less than 50%, and the activity-of the water Is

greatly reduced. The environment is no longer essentially aqueous. Up

to 507. acid, the absorbance peak remains the same, but the absorbance

increases by a factor of about 1.5. This change may indicate an ion-

association complex, such as Co^NO^ • The absorbance continues to

Increase as the concentration of sulfuric acid increases. The absorb­

ance maximum shifts to a higher wavelength, 550 nm, which is still in

the octahedral range. Probably a definite complex is formed with bisul­

fate ion in fairly concentrated sulfuric acid. In 70% perchloric acid

(1 1 .8 M) there is a smaller shift in the absorbance maximum to a higher

wavelength, 523 nm. However, it is also well known that the absorptivity

coefficients are a function of the index of refraction, which also changes

markedly in the concentrated acids.

Spectrophotometry in Hydrochloric Acid

The absorbance curve of cobalt(ll) in hydrochloric acid solu­

tions is the same as that in solutions of other inorganic acids up to a/ *concentration of about 5M. Above 6M the color changes to blue, and there

is strong absorption above 650 nm. Apparently a tetrahedral complex forms

in this concentration of acid since less energetic light (red) is necessary

to produce an electronic change. Figure 4 shows the abosrption curves

above 600 nm, the tetrahedral region. The greatest increase in absorbance

occurs when the acid is about 8M. All curves in the acid concentrations

above 6M have the same general shape, which indicates that only one

tetrahedral chloride complex is formed.

FIGURE A

ABSORBANCE OF Co(ll) IN CONCENTRATED HC1

(C_ s= 0.002F; HC1 molarity as indicated) Co

1.20

1.00(tO,8080

SB0.6077

0.40

0.20 6.6

750 650 600 nm700

Another interesting effect, shown in figure 5, is that the

absorbance at around 530 nm, the octahedral region, also increases as

the acid concentration increases up to about 8M. Thus an octahedral

complex, such as CoCl^cOg** or CoClCl^O),-+ , is still present in signifi­

cant amounts when the acid concentration is less than 8 M. Above this

acid concentration the octahedral complex appears to be changed to a

tetrahedral complex.

Another experiment involved changing the cobalt concentration

and keeping the hydrochloric acid concentration at 12M. The linear rela­

tionship of absorbance and cobalt concentration is shown in figure 6 .

The typical Beer’s Law curve shows that cobalt can be determined quanti­

tatively as well as qualitatively in concentrated hydrochloric acid.

Ion-association complexes containing more than one cobalt atom, such as

CoCl^CoCl^ , are ruled out because these would dissociate increasingly

during dilution, and Beer’s Law would not be followed. The sharpness of

the increase in absorbance with increasing acid concentration is shown in

figure 7.

4 Since electrophoresis and solvent extraction.experiments -

indicated the presence of a neutral complex, there was an effort to show

if the equilibrium constant is a function of the square of- the chloride

concentration. A modification of a method developed by Ramette (1963)

was used for the determination of a stability constant from absorbance

measurements. Since a one-centimeter cell was used, the term, "b," the

length of the light path, was omitted from the calculations.

FIGURE 5

ABSORBANCE OF Co(ll) IN 500-700 run REGION (C^o ss 0.10F; HC1 molarity as indicated)

1.40

1.20

1.00

0.80

0.60

040

0.20U

750

!0

600 550 500 nm

45

FIGURE 6

0.9

0.7

0.5

0.3

.2 .4 .6 .8 1.0 1.4 - 2M o l a r i t y , Co + + x I 0 “ 3

ABSORBANCE OF VARIOUS CONCENTRATIONS*OF COBALT in 12F HC1 at 694 nm

, 69

4mju

,FIGURE 7

1.0

0.8

<

0.4

0.2

1 0 IE2 84 6M o l a r i t y , HCI

ABSORBANCE OF 0.002M Co++ IN HCI OF VARYING CONCENTRATIONSat 694 nm

47

Assume that the only important complex is neutral CoCl2< The

hydrated cobalt(ll) ion does not absorb light at the wavelength used.

[GoCi 3h + ♦ ' — (1)2 [co++] [Cl y

From the basic equation of spectrophotometry, A » ebc, and fori

the conservation of cobalt and chlorine, one may write:

[CoCl2] = A/e (2)

CCo = tco++] + [CoCl2] (3)

ccl « 2[CoCl2] + [Cl-] ~ [bl-] (4)*

Since the total cobalt and chloride concentrations and the

absorbance are measured quantities, two unknowns, 0 2 and e, remain, which

can be determined graphically.

Substitute equivalent quantities from equations (2), (3), and

(4) into equation Cl),A/e

P, = ----------------- 2 • <5)(CCo - A/e)[cl”]

Multiply both sides by the denominator,

P2CCo[Cl" ] 2 - P2 [Cl“]2A/e = A/g (6 )

Multiply both sides by e/CCo[Cl” ] 2

A/C0 o[Cl- ] 2 = P2C - P2A/CCo (7)

48

Equation (7) is of the form, y = mx + b, in which y = c [c l- ] 2

b s= p2G > atlc* x “ A ^ C o * *r^e 8tability constant, P2 , equal to -m, or

the negative slope. Without knowing e, one can solve for P2 graphically

by plotting the values of x and y at different chloride concentrations.

The data are shown in table 7 and plotted in figure 8. The calculated_2value of $ 2 *-s " 1«33 x 10 , which is impossible. Something else is

happening besides the addition of chloride to cobalt ion to form CoCl2.

Since the activity of water is not constant in concentrated solutions,

the complete reaction must be considered:

C o (H20)6++ + 2C1“ ^ CoCl2(H20y2 + 4H20 (8)

Considering this effect, the real equilibrium expression should be

[CoC1,(H90)9] [h 9o ]a .p = “ T C9)2 [ C o C H ^ ) ^ ] [Cl"]2

T. E. Moore and his coworkers (Moore, Gootman, and Yates 1955, p. 298)

state that the activity of water in 9M hydrochloric acid is only 0.45

instead of 1.0. The activity coefficients of hydrochloric acid and

lithium chloride rise above 1.0 when the concentration is over 2m

(W. J. Moore 1962, p. 351). This combination of high cloride activity

drives the reaction (equation 8) to the right. No data could be found

for the activity of cobalt ion'at these high ionic strengths.

49

TABLE 7

ABSORBANCE DATA FOR P2 FOR CoCl2 IN HCI

CC1 CCo cci2 A694 x = A'CCoA

y - 2 CCoCCl

6.0 0.010 36 ‘ 0.214 21.5 0.595

6.6 0.004 43.7 0.226 56.7 1.294

7.2 0.004 52.0 0.505 126.3 2.43

7.45 0.004 55.7 0.685 171.5 3.07

7.8 0.004 61.0 0.93 233 3.81

8.15 0.004 66.5 1.28 321 4.81

8.4 0.002 70.8 0.73 365 5.15

9.0 0.002 81 0.98 490 6.05 '

9.6 0.002 92.3 1.06 -530 5.74

10.8 0.002 117 1.10 550 4.70

6

OJ _ o

o

o's.<

5

4

3

2

I 100 200 A / C

300 400'Co

GRAPH FOR P2 FOR CoCl2 IN HCI o

FIGURE 8

Spectrophotometry at High Ionic Strength

51

Another experiment was performed to observe if the blue tetra­

hedral complex forms in lower chloride concentrations when the ionic

strength is high. The molarity of both concentrated hydrochloric acid

and 707. perchloric acid is about 12. Thus the ionic strength can be

kept approximately constant by adding the desired amount of concentrated

hydrochloric acid, and then diluting to the required volume with 70%

perchloric acid. The blue color, as is shown in table 8, appears at a

much lower chloride concentration. With the high hydron,ium ion concen­

tration, there are few available water molecules left to form bonds with

the cobalt ions. Therefore, lower concentrations of chloride are needed .

to displace water from the hydrated ion and form the CoCl^ complex. The

dichloro complex is almost completely formed in 2M hydrochloric acid at

an ionic strength of 12. The stability constant,' approximately 6.

The chloride concentration is known only approximately because gaseous

HCI bubbles form and escape as soon as the perchloric acid is added.

The absorbance decreased about 257. after a solution that was initially

1.0M in HCI and 11M in HCIO^ stood for four hours in a closed volumetric

flask. Adding a small amount of water was found to eliminate the evolu­

tion of gaseous HCI. The absorbances at two ionic strengths, 8M and 10M,

are listed in tables 9 and 10. As long as the environment was essentially ■

perchloric acid, Ramette’s (1963) equation was found to hold. A plot of

the data in 10M solution is shown in figure 9. The calculated slopes are

-0.36 in the 8.0M solution and -2.64 in the 10.0M solution. Thus, at

25°C the corresponding values are 0.36 and 2.64, corresponding to an

• 52

TABLE 8

ABSORBANCE DATA FOR CoCl2 IN PERCHLORIC ACID AT [i = 12

(CCo a 0.002F)

Molarity, HCI A694

0.24 0.1570.36 0.241

0.48 0.515

0.60 0.5850.84 0.700

0.96 0.84

1.44 0.96

1.80 1.00*

increase of about sevenfold as the ionic strength is increased from 8 to

10. Thus, the lowered activity of the water plays a very important part'

in the reaction, when the ionic strength is so high. The absorbance in

8M acid reaches a maximum where the HCI contribution is about 3.6M.

Then the absorbance decreases somewhat when over 507. of the total acid is hydrochloric. -

Another experiment performed at high ionic strength involved the replacement of the hydrogen ion with lithium. The most soluble

available nonhalide salt of lithium was found to be the nitrate. By means

of gravimetric determination as the sulfate, the solubility of lithium nitrate was found to be 9.32M at 24.5°C. Then lithium chloride and nitrate

53

TABLE 9

CoCl2 IN PERCHLORIC ACID AT (i = 8.0

(C_ = 0.002GF)uO

HCI, M (CC1)2 A 694 A'CCoA

CCo(CCl)2

0.60 0.36 0.060 30.0 83.3

0.96 0.923 0.195 97.5 105.8

1.20 . 1.44 0.295 147.5 102.3

1.44 2.08 0.385 192.5 92.5

1.80 3.24 0.470 235 71.6

2.40 5.76 0.575 288 50.0

3.00 9.00 0.670- 335 37.2

3.60 13.00 0.740 370 28.4

4.20 - 0.705 - -6.00 - 0.615 - -

54

TABLE 10

ABSORBANCE DATA FOR CoClg IN PERCHLORIC ACID AT (-1 = 10.0 '

(CCo = 0.0020F)

HCI, M <cci)2 A694 A'CCo A'cc0cci)2

0.24 0.0576 0.101 50.5 877

0.48 0.231 0.515 257.5 1014

0.72 0.520 0.74 370 712

0.84 0.708 0.85 425 600

0.96 0.923 0.92 460 498

1.20 1.44 1.04 502 . • 349

1.56 2.44 1.18 590 * 242

1.92 3.70 1.17. 585 158

55

FIGURE 9

GRAPH FOR 02 FOR CoCl2 AT \1 = 10

800

600

400

4 200

600500400300

56

were mixed in proportions such that the total lithium ion concentration

was either 8 or 9M. In the 8M solution the molar absorptivity of cobalt

did not reach 6.0 at 694 nm until the chloride contribution was 3.0M.

Apparently the lithium ion is not as highly hydrated as the hydrogen

ion so the water activity is greater. The data for the 9M solution are

shown in table 11. Here the absorbances are somewhat greater, but the

slope is positive. The replacement of nitrate ion by chloride affects

the activity of water, and Ramette's equation does not hold.

TABLE 11

ABSORBANCE DATA FOR CoCl2 IN LiNO^ SOLUTION

(CCo = 0.0040M; |J = 9.0)

CC1 Ccci)2 A694- A 'CCo ‘ A 'CCo(0Cl>2

2.00 4.00 0.017 4.2 1.052.40 5.76 0.025 6.3 1.09

/ 3.00 9.00 0.054 13.5 1.503.60 13.00 0.085 21.3 1.644.20 17.65 0.129 32.3" . 1.834.80 23.1 0.193 48.3 2.096.00 36.0 0.437 109.3 3.04

57

Copper(II) in Hydrochloric Acid Solution

Like cobalt, copper also forms a complex in hydrochloric acid

having a color different from that of the hydrated ion. The absorbance

of the hydrated copper(II) ion occurs above 700 nm. This absorbance,

like that of the cobalt ion, increases slightly as any common acid is

added. By analogy with cobalt, it would appear that copper also forms

an ion association complex, such as CuCH^O)^ NO^ • More stable complex­

es must form in dilute solutions of chloride because then the absorbance

of copper(ll) is quite different in the ultraviolet, as shown in figure

10. At around 250 nm absorption is evident even in 0.10M chloride

solution. The absorbance increases with chloride concentration up to

12M, but not in a linear fashion. When the chloride concentration is ^

over 2M, the absorbance maximum begins to shift from 248 to 274 nm. At

the same time a new maximum begins to develop at 383 nm and a shoulder at

238 nm. From 7.8 to 9.6M chloride the absorbance at 274 nm is constant,

and the molar absorptivity is 4,500. Since there is no apparent movement

during electrolsis in 9M HCI, the neutral complex, CuCl^, is probably

present. The lack of a 383 nm peak and the shift of the 248 nm peak to

274 nm indicate that a lower chloride complex, probably CuCl+ , is

present when the chloride concentration is less than 2M. The high

absorptivity of copper chloride in the ultraviolet is generally assumed

to be due to charge-transfer.

When the chloride concentration is raised above 10M, subtle

changes occur in the absorbance curve. Perhaps the most significant is

that the absorption at 274 nm begins to rise again. The shoulder that

58

FIGURE 10

ABSORBANCE OF Cu(ll) IN HCI

(C„ = 0.0004F; HCI molarity as indicated)Cu

1.80

1.60

1.40

l7 81.20

1.00

Q80/

06078

0.40 10.8

0.20 6.0 4.83.6

450 350400 300 250 nm

was at 238 nm in 7.8M chloride shifts to 233 nm, and the absorbance in

this region decreases slightly. The absorbance at 383 nm continues to

increase in an almost linear fashion, and the wavelength of maximum

absorption shifts about four nanometers toward the ultraviolet. This

Indicates that a higher complex, probably CuCl^", is being formed at

these high chloride concentrations. The presence of a complex anion

in 12M hydrochloric acid was also confirmed by electrophoresis since

the spot was moving toward the anode when the paper disintegrated in

the concentrated acid.

More detailed data at the two maxima are listed in table 12.

These are enough to calculate (Ramette 1963) the first stability

constant: = 0.65. The absorbance at 248 nm was corrected for the

slight absorbance of copper(II) as perchlorate. ■

As with cobalt, either Kg or Pg for copper chloride could not

be determined unless the ionic strength was held constant by means of

an indifferent electrolyte. Perchloric acid was used to maintain the

ionic strength at 8.0, and the absorbance data are listed in table 13.

The calculated Kg from the slope is 0.64. Note that Kg is for the

reaction involving the addition of the second chloride to copper:

CuCl+ + Cl" CuClg.

The addition of the first chloride may be regarded as complete because

the absorptivity of copper in a solution, 0.12M in HCI and 7.88 M in

HCIO^, was greater than 2,000 at 248 nm. (See table 14.) At this

chloride concentration the absorbance at 383 nm was barely perceptible.t

The data for the calculation of for copper chloride in 8M perchloric acid are shown in table 14. At these low chloride ,

60

TABLE 12

ABSORBANCE OF COPPER(ll) IN HYDROCHLORIC ACID

(C„u = 0.0004M) Cu

HCI, M A248 A260 A274 A383

0.0 0.02 0.01 0.0 0.000.10 0.07 0.05 0.02 0.000.20 0.10 0.08 0.04 0.000.40 0.17 0.12 0.06 ' 0.000.60 0.22 . 0.17 0.08 0.000.80 0.26 0.20 0.10 0.001.00 0.31 0.25 0.13 0.001.80 0.46 0.41 0.25 0.002.40 0.61 0.57 0.40 0.003.60 0.81 0.85 0.72 0.024.80 0.93 1.05 1.06 0.076.0 0.98 1.19 1.38 0.216.6 1.00 1.25 1.52 0.277.2 - - 1.65 0.367.8. 1.06 1.43 1.78 0.468.4 0.99 1.35 1.77 0.529.0 - - 1.77 ‘ 0.569.6 0.97 1.38 1.78 0.6110.8 0.99 1.48 1.90 0.6811.9 1.00 1.54 1.93

'

0.76

TABLE 13

CuCl2 FOR K2 IN HC104

(C- = O.OOIO; |i = 8.0)Cu

HCI, M A o383 A 'CCu A'CCuCCl

0.36 0.293 293 8130.60 0.500 500 8330.96 0.705 705 7351.20 0.815 815 6791.44 0.895 895 6211.80 0.99 990 5502.40 1.10 1100 459

TABLE 14

DATA FOR K, FOR CuCl+ IN HC10 1: * 1(cCu = 0.0010; H = 8.0)

CC1 A248 Not A248 CC1 + CCu ACCC1 + CCu) A

CCuCCl CCuCCl

0 0.075 - - - -

' 0.002 0.175 0.100 0.003 150 . 5.00 x 10+0.004 0.293 0.218 0.005 273 5.450.006 0.402 0.327 0.007 382 5.450.008 0,560 0.485 0.009 545 6.060.010 0.638 0.563 0.011 619 5.630.014 0.738 0.663 0.015 711 4.740.016 0.788 0.713 0.017 756 4.450.020 0.880 0.805 0.021 844 4.020.12 >2.0 - - - -

concentrations, the concentration of copper ion is significant in

comparison. Development of the Ramette (1963) equation leads to the

terra, A(CC1+ CCu>/GCuccl» 'for x and A/CCuCcl for y. One must subtract 0.02 from the measured absorbance to correct for the absorbance of pure

copper perchlorate at 248 nm. From the slope one obtains a value of 73

for at 25°G for the reaction,

Cu** + Cl" - CuCl+ .

This stability constant increases as the temperature is raised. This

phenomenon was investigated only qualitatively. If one uses the value

of 73 for K^, he can calculate that conversion to CuCl+ is 987. complete

at a chloride concentration of 0.6M. It is at this concentration that

the straight line portion of the curve for begins, so one can assume

that CuCl+ is completely formed before C u C ^ starts to form.

Cobalt in Hydrobromic Acid Solution

The investigation continued with the study of cobalt(ll) in

concentrated hydrobromic acid. Qualitatively the reactions are -the

same as in hydrochloric acid. (See fig. 11.) As with the chloride,

there is only one absorbance curve other than that of the hydrated

cobalt ion. When the acid concentration is less than 6M, there is no.

measureable absorbance above 600 nm. This fact indicates that the

cobalt bromide complex Is less stable than the chloride complex, which

begins to show a blue color in ‘5M chloride solution. The shape of the

bromide absorbance curve is similar to that of the chloride, and there

are twin peaks at 696 and 723 nm.

63

FIGURE 11

ABSORBANCE OF Co(Il) IN HBr

(Cj, =s 0.002F; HBr molarity as indicated)

1,20

1.00 18.81

0.80

0.608,19

0.40

0.20705

700 650 nm

64

As with the other di-halide constants, where the significant

change occurs when the halide concentration is over 6M, the ionic

strength must be held constant preferably with perchloric acid. The

absorbance data obtained in 8.0M mixed hydrobromic and perchloric acids

are shown in table 15. The calculated slope is -0.25, and *-8 0-25.

Concentrated hydroiodlc acid is only 6M, and. at this concentra­

tion there was no visible evidence of a blue cobalt iodide complex.

TABLE 15

ABSORBANCE DATA FOR CoBr2 IN PERCHLORIC ACID

(CCq = 0.0020M; \1 = 8.0M)

CBr . CCBr)2 A723 A'CCo A^CCo(CBr)2

1.28 1.645 0.300 150 91.2

1.60 2.56 0.445 223 86.8

1.84 3.39 0.575 288 84.8

2.00 4.00 0.625 313 • 78.1

2.16 4.68 0.680 340 72.6

2.40 5.76 0.748 374 64.9

2.56 6.58 0.760 380 57.7

2.88 8.32 0.810 405 48.6

Summary

65

With electrolyte concentrations greater than 6M, as in

concentrated acids, the activity of water becomes considerably less, and

its activity cannot be considered to be constant.. The difference in

hydration of two "indifferent” ions may cause a large difference in the

apparent stability constant.The apparent constants at 25°C, summarized in table 16, are in

.effect conditional constants.

' TABLE 16

STABILITY CONSTANTS IN AQUEOUS SYSTEMS

Complex Ionic Strength Constant

CuCl+ 0.5 to 2.0 ^ = 0.65

CuCl+ 8.0. Kx = 734

CuCl2 8.0 K 2 = 0.64

CoCl2 8.0 P2 = 0.36

CoCl2 10.0 02 = 2.64

CoBr2 8.0 02 = 0 . 2 5

66

From spectrophotometry alone the highest and only cobalt

chloride complex found in aqueous solution appears to be neutral

CoC^CHgO^- Electrical migration, electrophoresis, and solvent

extraction experiments to be discussed in chapter four suggest that

much of the cobalt is present in concentrated hydrochloric acid- as at

neutral complex. However, the electrical migration and ion exchange

experiments hint that both positive and negative 6pecies are in

equilibrium with the neutral complex.

CHAPTER IV

COBALT(II) CHLORIDE COMPLEXES IN ORGANIC SOLVENTS

Electrical migration experiments suggested but did not confirm

that the principal cobalt(ll) chloride complex in concentrated hydro­

chloric acid in neutral CoCl^* Often solvent extraction can indicate

whether a metal complex is neutral or a charged ion. However, the rule

Is not infallible since an ion-association complex, which is neutral,

can be formed and is extracted by an organic solvent. Thus in the ether

extraction of iron(III) chloride, it is (C^H^JgOH^eCl^- that is

extracted.

Simple solubility experiments were tried first. Cobalt(ll)

chloride was found to be insoluble in nonpolar solvents such as benzene

and chloroform. The solution in ethyl ether has a faint blue color and-5the molar solubility is about 6 x 10 if one assumes the same absorp­

tivity as in acetone in which e = 200. In ketones and alcohols, cobalt

and copper(ll) chlorides are very soluble. Except for methanol in

which a solution of cobalt chloride is pink, all of these solvents

produce a blue solution. Since no excess chloride was added, the

solution probably does not contain anionic complexes, such as CoCl^ in any high concentrations.

67

68

Solvent Extraction

Alcohols and ketones can extract cobalt and copper(II)

. chlorides from 12F lithium chloride and hydrochloric acid. Extraction

experiments were performed using methyl isobutyl ketone and n-heptanol

as the organic phase.

Metfhyl isobutyl ketone was completely miscible in 12F hydro­

chloric acid but when the acidity was reduced to 9F, two layers appeared.

In the presence of cobalt(ll) both phases were blue in color. The

absorpion curve in the red end of the spectrum, above 600 nm, was nearly

the same for both layers. The ketone layer also showed considerable

absorption below 500 nm. This latter absorption occurs even when no

pobalt is present. Since the color deepens upon standing, the reaction

may be oxidation or polymerization catalyzed by hydrogen ions.

N-heptanol formed a separate phase in contact with all concen­

trations of hydrochloric acid. Fortunately there were no side organic

reactions. The same absorption peaks were found in both layers, indicat­

ing that the same chloride complex is present in the aqueous and organic

layers. As shown in figure 12, the alcohol extracted about half of the-

cobalt chloride from either 12F HG1 or LiCl. When the acidity was

reduced to 8F, the amount of extraction was reduced to about half but the

shape of the absorption curve of the organic layer remained the same.

The absorption curve for the aqueous layer was different in one respect:

There was considerable absorption at 530 nm, indicating that part of the

cobalt is octahedral, presumably C o C ^ O ) ^ * . From 6F hydrochloric acid

there is little extraction indicating that nearly all the cobalt is

FIGURE 12

EXTRACTION OF COBALT CHLORIDE WITH N-HEPTANOL FROM HC1

(CoCl2 = 0.00033F)

1.00

0.80

0.60

0.40

0.20

740 700 660 620 nm

70

present as the hydrated ion. The blue extractable complex could be

neutral CoCl^Csol v e n t ) o r an ion-association complex such as

H30+CoC13".

Properties of Cobalt(ll) Chloride In Acetone

Even with no excess of chloride, anhydrous cobalt chloride

dissolves in acetone, forming a dark blue colored solution. As stated

on page 8, Fine (1962) found that three chlorides add almost quantita­

tively to form the tetrahedral complexes CoCl^ and then CoCl3” . He

listed equilibrium constants for their formation, but gave insufficient

data for the calculations. Cobalt nitrate and perchlorate are also

soluble in acetone, but they produce pink complexes which presumably

are octahedral. Since it was found that cobalt(ll) forms a nitrate

complex in acetone, perchlorates were used in preparing all standard

cobalt solutions. These were mixed with acetone solutions of lithium

chloride so that the mole Cl:Co ratio varied from zero to 100.

Cobalt perchlorate has the same spectrum in acetone as in

yrater. As the chloride concentration was increased, four different

types of absorbance curves were obtained, indicating that there are

possibly four different cobalt complexes, Co(acetone)g++, CoClg, CoCl'3~, - 2and CoCl^ . The last three are blue, and presumably, tetrahedral. A

family of curves, in which the Cl:Co ratio varies from zero to sixty, is

shown in figure 13. The variation in absorbance at selected wavelengths

is shown in figure 14.

At any wavelength in tfre red region, there appears to be a

linear increase in absorbance as the chloride ratio increases from zero

71

FIGURE 13

ABSORBANCE SPECTRUM OF Co(Il) IN ACETONE WITH ADDED LiCl

<CCG « 0.003F)

Cl:Co Ratio is Indicated

1.00 - I

V\0/

V /0.80-

0.60-

0.40

0,20

740 700 660 620 580 nm

V72

FIGURE 14

ABSORBANCE OF Co(ll) IN ACETONE WITH ADDED Cl" AT SELECTED WAVELENGTHS(C„ = 0.003F)Co

=* * £ B B E° o 2O r o O CD CD r**

< O □

oa:

O

<3 '

■ '• ■' 73

to two. The absorption at 520 run also disappears as the two chlorides

are added. This evidence indicates that no appreciable concentration

of the CoCl+. complex is present at equilibrium. Since the complex is

probably tetrahedral, the following reaction probably occurs:

c0(c3h6o) 6++ + 2 c i" -* co(c3h6o) 2c i 2 + 4c3h6o

The addition of two chlorine atoms to cobalt was shown to be essentially

quantitative. At constant cobalt concentration the absorbance is a

linear function of the chloride concentration. This was the basis for a

new method for the quantitative determination of chloride, which will be

described in detail in chapter five.

After the first two chlorine atoms have been added, a third is

added almost quantitatively, and the absorbance curve changes. The most

prominent feature of the absorbance curve for 0 0 0 1 2 " is a deep minimum

at 620 nm. This minimum disappears as still more chloride is added but

the addition of the fourth chloride is not quantitative. Not until the

mole Cl:Co ratio is 30:1 does the absorbance become essentially constant.

Use of Spectrophotometry to Determine Stability/ Constants by the Slope-Intercept Method

The final stability constant, K^, which refers to the reaction, CoCl3“ +. Cl" - CoCl4"2

may be determined from absorbance measurements by the slope-intercept

method (Whiteker and Davidson 1953). The slope-intercept formula has

two basic forms, when the equilibrium involves the addition of only one

ligand. * .

If the addition of the first ligand is considered,the equation

has the following form:

[L] /

It can be solved graphically. For the addition of the final ligand, the

following form is often more useful.

x - X - i + (H ' x 1 M K n < n )

Equation (11), was used in the present study. The measured

absorbance, A, is the sum of the absorbances of the two main components,

and

S - eMLN_1b[MLN-l] + = EbCH Cl2)

where 6 indicates the molar absorbance coefficient; b is the length of the light path in centimeters; concentrations are. expressed in moles per

liter, and is the total concentration of metal ion in any form.

The concentrations are related by the stability constant K.

= --- 3---- (13)

A = SbCM = 5b[MLN_1] + eb[MLN_1]KN (14)

ebCML,,^] + eCML^^CL]!^ . e b C M L ^ ] + (15)N-l

Cancel out the terms b and [ML^ ^3 and rearrange.

• 75

Since A = one can obta*-n equation (11), and plot the mea­

sured absorbance as y, and (A^ ^ - A)[L] as x. The intercept is A ^

and the slope gives the stability constant, K^. In the acetone solutions

of cobalt halides, it will be evident that N = 4. One obtains A ^ by4

saturating the solution with the lithium halide. The free halide

concentration is first estimated by subtracting the halide needed to

form ML^ from the total halide concentration. Then a rough calculation

is made, and a further correction made for the halide needed to form ML,.4The data for the calculation of for the tetrachloro complex

in acetone are shown in table 17, and the curve in figure 15. The slope,

which is equal to K^, is calculated to be 215.

TABLE 17

DATA FOR K, FOR CoCl “2 IN ACETONE 4 4

(Determine at 625 nm; CCq = 0.003M; A ^ = 1«00)4

CC1 ^625 ^ 4 APreliminary

FML4Corrected

[L] (am l 4"A)CL] [L] (Am -A>[L] • 4

0.036 0.87 0.13 0.027 0.0035 0.78 0.0247. 0.0032 ‘

0.018 0.76 0.24' 0.009 0.0022 0.59 0.0072 0.00174

0.0156 0.72 0.28 0.0066 0.00185 0.51 0.0051 0.00143

0.0120 0.58 0.42 0.0030 0.00126 0.29 0.0021 0.00088

0.0108 0.53 0.47 0.0018 0.00084 0.20 0.0012 0.00056

0.0096 0.44 0.56 0.0006 • 0.00034 0.05 • 0.0005 0.00028

FIGURE 15

K, FOR CoCl "2 IN ACETONE 4 4

0.9

0.7

0.5

0.3

ML.-S)U

77

Determination of Stability Constants by Means of Corresponding Solutions

The stabilities of CoCl^ and CoCl^” are too great to allow the

calculations of P2 and by means of the slope-intercept method. For

these calculations one may use Bjerrum’s method of corresponding solu­

tions (1944).By definition, corresponding solutions are those in which the

ratio of all complex species is constant. An example is a series of

solutions in which the ratio of CoCl2 to CoCl^ is constant. Then n as

defined below is constant.complexed ligand conc. - free L

n = total metal conc. (17)

If only mononuclear complexes are formed between the metal ion and the

ligand, it may be shown that the proportions of the various species and

the mean number of bound ligands per metal ion, n, depend only on the

ligand concentration. Thus, the fraction of the metal present as any

species, such CoCl^ in acetone, is[CoCl2]

2 [Co'1"r] + [CoCl2] + [CoCl3~] + [ClCl4_t]aCoCl0 “ ++-. • - • -2- • (18)

Each concentration may be replaced by the appropriate stability term,.

o __________________ P ^ C o ^ H C !-]2__________

0cC12 " [Co++] + P2[Co++] [ c r ] 2 + P2[Co++][Cl"]3 + P4[Co++][Cl"]4 (I9)

The term, CCo++] or the more general C*G, can be cancelled throughout equation (19), and it simplifies to that found in equation (20).

Therefore, It has been shown that the fraction of CoC]^ or any other

species is dependent only on the concentration of chloride ion or other

ligand, and is independent of total metal ion concentration. If poly-

nuclear species, such as 1*2^ 2 ’ are present, the free metal ion concentra­

tion does not cancel out, and .the right side of equation (2 0 ) cannot be

developed without an [M] term remaining. Fortunately, in the case of the

cobalt halides, polynuclear complexes do not appear to be formed.If the complexes absorb light, and the absorbances of solutions

having identical free ligand concentrations are measured in cells of such

length that the product of the cell length and total metal ion concentra­

tion are equal, they will yield identical absorbance curves, provided no

polynuclear species are formed. For instance, a 0.10M solution of a given

species in a 1 .0 -cm cell would have the same absorbance as a 0 .0 1M solu­

tion in a 10-cm cell. If it is not convenient or possible to use a cell

of the proper length, the absorbance readings can be normalized by multi- / ‘plying every actual absorbance by the factor, C /C , where the subscriptn &"n" refers to the concentration of the normal or standard solution and

the subscript "a" indicates the actual concentration. If the cell lengths

are different, an additional factor b /b is needed.n aConversely, if solutions of the same complex system have iden­

tical normalized absorbance curves, they contain the same free ligand

concentration. If this is true, then all the complexes must be mononu­

clear. By chance in the course of experimental work, two solutions of co­

balt iodides in acetone did produce nearly identical normalized aborbance

curves. In figure 16, each absorbance on curve B may be multiplied by the

factor 1.32, and curve A is obtained. The ratio of the principal species,

C 0 I2 and Col^*", and the concentration of free iodide appear to be 1:he same

in both solutions if the above principles are true. Bjerrum (1944) applied

the name "corresponding solutions,11 to solutions which yield the same

normalized absorbance curves and he applied his principles to the calcula­

tion of free [L] and n from absorbances and total concentrations.

According to the author’s preceptor, corresponding solutions may

be compared to solutions of a composite dye prepared by mixing pure dyes

in a definite proportion. The color tone remains the same even when the

solution is diluted. The color of two different dilutions of this mixture

can be matched by varying the length of the light path, just as is done In

visual colorimeters of the Dubosq type.

In a mixture of complexes the expression for the absorbance, A, in a cell of length, b, is _

A « ecMb = b([M ]e0 + [ML]er + [ML2 ]e2 + .......) (2 1 )

Here e is the mean molar absorptivity and e^, e^, and e2 are the molar

absorptivities of the species [M], [ML], and[ML2], Substituting the

product, Pi[M][L]i, for the concentration of each complex and factoring out [M] yields

A / M = ecMb/[M ] = b(eQ + el P1[L ] + e ^ L ] 2 ...............) (22)

One may write equation (20) as the fraction of metal present as [M] and

substitute into equation (2 2 )

80

FIGURE 16

ABSORBANCE CURVES OF TWO CORRESPONDING SOLUTIONS IN COBALT-IODIDE SYSTEM

0.80

0.60 —

0.40

0.20

f \

\

\\\

_ _ II___!_____ I760 720 680 640 nm

aCCo = 1.32 x 1Q”^F; C . = 3.6 x 10”^F in 1.0 cm cell,

= 1.0 x 10”^F; Cj. = 3.6 x 10”^F in 10 cm cell.

81

Thus, the mean molar absorptivity, e, of an equilibrium mixture of

mononuclear complexes is shown to be a function of the single variable [L],

or free halide ion, and is independent of the total concentration, C^. In

a particular system at a given wavelength, the e and P values are all

constant. At constant [L], e is constant. If the absorptivities of two

solutions at given wavelengths are equal, it follows that the ligand

concentrations of these solutions are equal. Conversely, if the free

ligand concentrations in the solutions are equal, their mean molar absorp­

tivities must also be equal.

The simplest way to recognize corresponding solutions is to plot

the normalized absorbance, or its equivalent e, of particular total metal

Ion concentrations at a particular wavelength against the ratio of total

ligand to metal ion concentrations. Several wavelengths should be chosen

which are near or at the maximum absorbance for particular concentration

ratios. Horizontal tie lines which intersect rising or falling portions

of the curves, indicate the compositions of corresponding solutions. An

example is shown in figure 17. The normalized absorbance curves are plotted

<for two concentration ranges, which differ by a factor of ten. Due to

slight dissociation, the absorbances of the solutions diluted tenfold in

a ten-cm cell are a trifle lower. Horizontal tie lines connect as many

pairs of corresponding solutions as one wishes.

In order to derive the mathematical relations between two

corresponding solutions, it helps to indicate the two different concentra­

tions by the subscripts, 1 and 2. In two corresponding solutions the

concentration of free ligand is the same. Furthermore, it is evident from

Bjerrum1s formation function that the mean number of bound ligands per

(Absorbance

at 680

nm normalized

to C_

as 0.006F)

82

FIGURE 17

SERIES OF CORRESPONDING SOLUTIONS FOR COBALT CHLORIDE IN ACETONE

O CCo=.004F A CQ = -006F

□ C, =.0004F Co1.6

1.2

1.0

0.8

0.6

0.4

0.2 0.6 1.0 1.4 1.8 Cl/co

metal Ion, n, Is the same for both solutions, since it.depends only on

[L] and the complex equilibrium constants. From the definition of n

(see equation 17), one may write the following for two correspondingsolutions M _ [L]

5 = ---- = (24)Ml M2

By eliminating ii and solving for [L], one obtains

[L] - ^ (25)Ml " M2

Hence, it is possible to solve for [L] in these two corresponding solu­tions. This value of [L] may be substituted in either of the n equations

(24), and a value of n can also be calculated. The method is roost satis­

factory when the differences between the two and two terns are

large. When the differences are small, the errors in reading the absor­

bance or neglecting a cell blank can become so serious that the results

are meaningless.Any of a variety of methods can be used for the calculation of

the equilibrium constants after values are obtained for [L] and n. The /

simplest, which is sati.sfactory for most purposes, is* to plot n against pL.

Then one assumes that the‘individual constants, K^, Kg, K^, etc., are equal

to the corresponding [L] values at n = 1/2, 3/2, 5/2, etc. There is an

analogy for a weak acid when it is half titrated, that is n for H+ is

equal to 1/2, that pK = pH. The method fails when the •equilibria

overlap. The extreme example occurs in the cobalt chloride system in

acetone, where the first two chlorides appear to add simultaneously or

not at all. Individual values for and Kg are meaningless, and 0g is- 2calculated as 1/[C1 ] at n = 1.0. For cases of moderate overlap,

84

Carlson, McReynolds, and Verhoek (1945) have developed an iterative solution

of.the constants from the exact Bjerrum formation functions at half-integral

values of n. Mrs. Weed (1964) used Bjerrum1s method to determine the first

three stability constants in the aqueous palladium(ll) chloride system.

The data for the calculation of P2 ^or the c°kalt chloride system

in acetone are shown at four wavelengths in tables 18'to 21. Figure 17

shows a plot of the data at 680 nm. A sample calculation for free chloride

ion concentration in one pair of corresponding solutions starts with

equation (25).The starting reference point is taken where C = 0.0004F and C , = 0.00048UO bland the normalized absorbance is 0.947. This point from solution 2 is

joined horizontally to the normalized absorbance curve of the more concen­

trated solution 1^ At the same absorbance, 0.947, the apparent concentra­

tion ratio is 1.145. Since the chloride concentration was constant in one

series at 0.006F, the cobalt concentration in the corresponding solution -

is 0.006/1.145 or 0.00524F. Now one can substitute these values into

equation ( 25) •0.00524(0.00048) - 0.0004(0.006) „ „ ,rt-5 .

LC 1 ] = --------- 0.00524 - 6:0004-'------ = ? *3 x 1 0 * (26)

- CC1 " £C 1 ' 0.0060 - 2.3 x 10" 5 . ■n e — -------= -----0 7 0 0 5 2 4 ----------- 1-14 (27)Co

Other values for [Cl“] obtained at 680 nm near ii = 1.0 are 2.1, 1.7, and-5 -51.8 x 10 . An average value is 2.0 x 10 .

P2 = 1/[C1" ] 2 = 1/4.0 x 10- 1 0 = 2.5 x 109 (28)Similar calculations were made for at wavelengths. The

9results are 2.8 x 10 at 660 and 670 nm, and 2.2 at 690 nm. The averageg

$ 2 value for C o C ^ is 2 . 6 x 1 0 .

85

TABLE 18

NORMALIZED ABSORBANCE DATA FOR CoCl2 AT 660 nm

(Normalized’ to C = 0.0060F)

CCo CC1 A(1.0 cm) A(10 cm) An

0.0030 0.0060 0.772 • • • • • 1.5440.0040 0.0060 0.754 • • • • • 1.1300.0050 0.0060 0.737 • • • • • 0.885

0.0060 0.0060 0.730 0.730• 0.0080 0.0060 0.714 .0.535

0.0120 0.0060 0.690 • • • ■ • 0.3570.0040 0.0072 0.905 • • • • « 1.3420.0040 0.0060 0.757 t • * ■ • 1.1350.0040 0.0048 0.595 « • • t • 0.8920.0040 0.0036 0.438 • * • * * 0.6570.0040 0.0030 0.355 • • « • m 0.533 -0.00040 0.00072 0.913 • 1.3700.00040 0.00060 0.721 1.082

0.00040 0.00048 0.570 0.855

0.00040 0.00036 0.410 0.615

0.00040 0.00030 0.337 0.506

86

TABLE 19 ,

NORMALIZED ABSORBANCE DATA FOR CoCl2 AT 670 tun

(Normalized to C_ = 0.0060F)r*

CCo CC 1 A(1.0 cm) A(10 cm) A . n

0.0030 0.0060 0.865 1.730

0.0040 0.0060 0.842 1.262

0.0050 0.0060 0.817 0.983

0.0060 0.0060 0.806 0.806

0.0080 0.0060 0.790 0.593

0 . 0 1 2 0 0.0060 0.757 0.379

0.0040 0.0072 1 . 0 2 2 1.533

0.0040 0.0060 0.845 • • • • • 1.267

0.0040 0.0048 0.665 1 . 0 0 0

0.0040 0.0036 0.493 0.740

' 0.0040 0.0030 0.402 • • • m • 0.603

0.00040 0.00072 1.015 1.523

0.00040 0.00060 . 0.826 1.239

0.00040 0.00048 0.653 0.980

0.00040 0.00036 0.465 0.698

0.00040 0.00030 0.383 0.575

87

TABLE 20

NORMALIZED ABSORBANCE DATA FOR CoCl2 AT 680 nm

(Normalized to Cn = 0.00060F)

CCo CC1 A(1.0 cm) A(10 cm) A . n

0.0030 0.0060 0.875 1.750

0.0040 0.0060 0.845 1.267

0.0050 0.0060 0.8.25 0.990

0.0060 .0.0060 0.805 0.805

0.0080 0.0060 0.790 0.593

0 . 0 1 2 0 0.0060 0.750 0.375

0.0040 0.0072 1.040 1.560

0.0040 0.0060 0.845 1.267

0.0040 0.0048 0.665 1 . 0 0 0

0.0040 0.0036 0.493 0.740

0.0040 0.0030 0.398 • * * • • 0.597

0.00040 0.00072 1.015 1.522

0.00040 0.00060 0.820 1.230

0.00040 0.00048 0.635 0.952

0.00040 0.00036 0.460 0.690

0.00040 0.00030 0.377 0.565

88

TABLE 21

NORMALIZED ABSORBANCE DATA FOR CoCl2 AT 690 nm

(Normalized to = 0.0060F)

CCo CC 1 A(1.0 cm) A(10 cm) An

0.0030 0.0060 0.740 1.480

0.0040 0.0060 0.698 1.047

0.0050 0.0060 0.667 0.800

0.0060 0.0060 0.662 0.662

0.0080 0.0060 0.647 0.486

0 . 0 1 2 0 0.0060 0.617 0.308

0.0040 0.0072 0.858 1.287

0.0040 0.0060 0.695 1.042

0.0040 0.0048 0.546 0.820

0.0040 0.0036 0.407 0.610

0.0040 0.0030 0.325 ..... 0.503

0.00040 0.00072 0.820 1.230

0.00040 0.00060 0.658 0.987

0.00040 0.00048 0.510 0.765

0.00040 0.00036 0.361 0.541

0.00040 0.00030 0.299 .0.449

89

At a given wavelength and cobalt concentration, the absorbance

curve is a straight line up to the point where the Cl:Co ratio is at

least 1.5. Then the slope increases, and this fact may indicate that

CoCl^” , as well as CoCl^, is beginning to form. For the entire plot of

absorbance at constant cobalt concentration versus the Cl:Co ratio at

700 nm, see figure 14.

The stability constant, for CoCl^” , can be determined in the

region where n = 2.5. Absorbance data at a wavelength on either side of

• the maxima for CoClg and CoCl^- appear in tables 22 and 23. The straight

line portion of a normalized absorbance plot occurs between the mole

Cl:Co ratios of 2.3 and 2.8. An isosbestic point at 682 nm indicates that

only two complexes are present when the ratios are between these figures.

The average value for [Cl~3, calculated for several correspond­

ing solutions in this region is 6.5 x 10”^. The value of n, where the

ratio is 2.5, is 2.47. This is one case where tbe values of n and C, :C,.L Mare nearly the same. The stability constant, K^, which is 1/[C1*"] at

n = 2.5, is calculated to be 1.5 x 1CT*.

An interesting fact is that the absorption spectrum of the

tetrachloro complex in acetone is practically the same as that of cobalt

chloride in concentrated hydrochloric acid. These spectra, along with

that of cobalt chloride in ethanol containing excess chloride, are shown

in figure 18. The similarity of the spectra of cobalt tetrachlorides

such as solid Cs2CoCl^, Li2CoCi^ in acetone, and C o C ^ in concentrated

HC1 has led many researchers to believe that all exist as tetrachloro

complexes. Yet the spectrum of cobalt chloride in ethanol containing

90

TABLE 22

NORMALIZED ABSORBANCE DATA FOR CoCl3 AT 674 nm

(Normalized to C„ = 0.0060F)1 oo

CCo CC1 A(1.0 cm) A(10 cm) An

0 . 0 0 2 0 0.0060 0.795 2.385

0 . 0 0 2 2 0.0060 .0.838 2.285

0.0024 0.0060 0.853 2.130

0.0026 0.0060 0.895 2.065

0.0028 0.0060 0.895 1.915

0 . 0 0 0 2 0 0.00060 0.795 2.385

0 . 0 0 0 2 2 0.00060 ■0.830 2.265

0.00024 0.00060 0.843 2.107

0.00026 0.00060 0.890 2.055

0.00028 0.00060 0 . 8 8 8 1.902

91

TABLE 23

NORMALIZED ABSORBANCE DATA FOR CoCl3“ AT 694 ran

t (Normalized to = 0.0060F)

CCo CC1 ACl.O cm) A(10 cm) An

0 . 0 0 2 0 0.0060 0.890 2.67

0 . 0 0 2 2 0.0060 0.853 2.325

0.0024 0.0060 0.810 2.025

0.0026 0.0060 0.742 1.712

0.0028 0.0060 0.692 1.480

0 . 0 0 0 2 0 0.00060 0.872 2.62

0 . 0 0 0 2 2 0.00060 0.850 2.320

0.00024 0.00060 0.805 2.015

0.00026 0.00060 0.735 1.700

0.00028 0.00060 0.683 • 1.460

FIGURE 18

92

ABSORBANCE OF COBALT CHLORIDE IN VARIOUS SOLVENTS(C_ = 0.002F)

1.20

1.00

0.80

0.60

0.40/

0.20 — /

740 700 660 620 580 nmAcetone with 0.1F LiCl Ethanol with 0.1F LiCl Aqueous HC1, 12F

93

0.10M chloride is like that of the dichloride in acetone. Ethanol

solutions are discussed more thoroughly later in this chapter., .

Effect of Water in the Acetone

The fact that traces of water in the acetone may affect the

absorbance was apparently not investigated by Fine (1962). The acetone

used in this laboratory,which was purchased from J. T. Baker Chemical

Company, contained 0.337. water. In order to avoid further contamina­

tion, anhydrous salts were used in all quantitative work. To determine

if water had an appreciable effect, up to 17. by volume of additional

water was added. Then spectrophotometric measurements were made where

Chloride to cobalt ratio was 2:1 and 8:1. For both ratios the only

effect was to move the absorbance maximum about five nanometers toward

shorter wavelengths. The molar absorptivity at the' maximum remained the

same. When the added water content was 1.57., there was visual evidence

that the absorbance in the red had decreased. The conclusion is that

added water up to 0.507. does not affect the absorbance measurements

significantly, and reagent-grade acetone may be used.

Spectra of Cobalt Bromide Solutions

The spectra of acetone solutions of cobalt(ll) containing

lithium bromide are shown in figure 19. The spectra are similar to those

of the corresponding chlorides except that the dibromide has only a

single absorption peak. The tri- and tetra-bromides also appear to be

somewhat less stable than the corresponding chlorides. The absorptivities

of all three bromide complexes are considerable higher than those of the

94

FIGURE 19

ABSORBANCE OF Co(ll) IN ACETONE WITH ADDED LITHIUM BROMIDE(Co++ = 0.002F)

Ratio of Bromide to Cobalt as Indicated

1.40-

1.00 -

0.60

0.20

\//// ,

740 700 660 620 580 nm

95

corresponding chlorides, and the absorption peaks of the bromides appear

at a somewhat higher wavelength.The calculations of the stability constants for the bromides

can be performed in the same way as for the chlorides. The same slope-

intercept equation (11) (Whiteker and Davidson 1953) can be used for the

determination.- of the final constant K^, which was found to be 65. The

absorbance data appear in table 24. Again the concentration of free

bromide is corrected first for the ligand necessary to form ML^, and

then for the additional bromide used to form ML^. The intercept, which

is , is 1 .0 1 .

TABLE 24

ABSORBANCE DATA FOR K. OF CoBr.**2 IN ACETONE AT 715 nm4 4 ^(CM = 0.002F; Am l = 1.70)

CBr A v z4

PreliminaryFML.4

Corrected[L]

4

1-1

1i_i

;

(ah l -A)[L] 4

,0.062 1.55 0.15 0.056 0.0084 0.78 0.0545 0.0082

0.031 1.44 0.26 0.025 0.0063 0.62 0.0238 0.0062

0.0188 1.32 0.38 0.0128 0.0049 0.45 0.0119 0.0045

0.0156 1.24 0.46 0.0096 0.0044 0.33 0.0089 0.0041

0.0125 1.18 0.52 0.0065 0.0034 0.25 0.0060 0.0031

0.0094 1.13 0.57 0.0034 0.0019 0.17 0.0031 0.0018

96

The constants, P2 ant* ^3 * can be determined as in the chloride

system by means of Bjerrum's corresponding solutions (1944). The absorb­

ance data for the calculation of fL at two wavelengths appear in tables 25u ^

and 26. A graph showing the normalized absorbance curves at 670 nm is

shown in figure 20. When the solutions were diluted tenfold, the absorb­

ances were measures in a ten-centimeter cell.

TABLE 25

NORMALIZED ABSORBANCE DATA FOR CoBr2 AT 670 nm

(Normalized to C_ = 0.0050F)to

000

CC1 A(1.0 cm) A(10 cm) An

0.0030 0.0060 1.090 • 1.8170.0040 • 0.0060 1.118 1.400

0.0050 0.0060 1.138 1.400

0.0060 0.0060 1.130 1.138

0.0080 0.0060 1.117 0.945

' 0 . 0 1 2 0 0.0060 1 . 1 0 0 ..... 0.458

0.00030 0.00060 • 1.145 1.910

0.00040 0.00060 1.155 1.445

0.00050 0.00060 1.128 1.128

0.00060 0.00060 1.113 0.928

0.00080 0.00060 1.080 0.675

0 . 0 0 1 2 0 0.00060 1.019 0.425

97

TABLE 26

NORMALIZED ABSORBANCE DATA FOR CoBr2 AT 680 nm

(Normalized to C^o = 0.0050F)

..... CCo CBr A(1.0 cm) A(10 cm) A. n .....

0.0030 0.0060 1.145 1.910 .

0.0040 0.0060 1.142 1.430

0.0050 0.0060 1.155 1.155

0.0060 0.0060 1.148 0.957

0.0080 0.0060 1.133 0.708

0.0120 0.0060 1.113 0.464

0.00030 0.00060 .•••** 1.165 1.933

0.00040 0.00060 1.142 1.430

0.00050 0.00060 1.115 1.115

0.00060 0.00060 1.090 0.910

0.00080 0.00060 1.050 0.656

0.00120 0.00060 0.980 0.408

0.6 10

1-4 Br/Co

CORRESPONDING SOLUTIONS FOR CoBr2 IN ACETONE AT 670 nm

(Normalized to C* = 0.0050) oo

. ■ . 99

With a close inspection of the curves, one can see that there

is a slight bend at the point when the molar Br:Co ratio is about 1.2.

The absorbance maximum shifts from 678 nm, when the ratio is 2.0‘, to*4*673 nm, when the ratio is 0.5. These are two hints that a CoBr complex

forms first, and is stable in small concentrations when there is insuffi­

cient bromide to form CoBrg. Furthermore, at wavelengths below the

maximum absorbance, the molar absorptivity in terms of total cobalt is

actually greater in the more dilute solution, if the ratio of bromide to

cobalt is 1.5 or more. In the more dilute solutions one would expect to

find a higher relative concentration of CoBr+ , which apparently has the

greater absorptivity at wavelengths below 675 nm.

No attempt was made to determine quantitatively the individual

stability constants, and Kg. Calculations show that the apparent

free bromide concentration is lower if the Br:Co ratio is 1.5 than when

the ratio is 0.5. This indicates that Kg is higher than and that

CoBrg is formed from CoBr+ before all the free Co++ has reacted. The

equilibrium may be expressed as follows:

t Co++ + 2Br” 51 CoBr+ + Br~ ** CoBrg.

For the determination of the equilibrium constant, Pg,

corresponding solutions, in which the Br:Co ratio varied-from 0.5 to 1.5,

were solved for free [Br""] by means of equation (25). The average of

about 20 [Br ] values, an equal number on either side of the 1:1 ratio_5and on either side of the maximum absorbance peak, was 2.5 x 10 . The

- 2 9value of Pg is found to be l/[Br ] or 1.6 x 10 .

100

The calculation of for the formation of CoBr^ is straight­

forward because only CoBr2 and CoBr^ appear to 'be present between B n C o

ratios of 2.2 and 3.0. An isosbestic at 6 8 6 nm tends to confirm that

only two complexes are present in this ratio range. The data for the

calculation of at two wavelengths are shown in tables 27 and 28.

TABLE 27

NORMALIZED ABSORBANCE DATA FOR CoBr3” AT 680 nm

(Normalized to C„ = 0.0030F)Vj O

CCo Gfer A(1.0 cm) A(10 cm) An

2 . 0 x 1 0 " 3 6 . 0 x 1 0 " 3 1.025 1.5372 . 2 6 . 0 1.070 1.4602.3 6 , 0 1.090 1.424

2.4 6 . 0 1.097 1.365

2 . 6 6 . 0 1.133 1.307

2.7 6 . 0 1.150 1.280

3.0 6 . 0 1.148 • 1.148

2 . 0 x 1 0 " 4 6 . 0 x 1 0 ” 4 1.025 1.537

2 . 2 6 . 0 1.035 .1.410

2.3 6 . 0 1.083 1.414

2.4 6 . 0 1.075 1.340

2 . 6 6 . 0 1.118 1.290

2.7 6 . 0 1.150 1.280

3.0 ■ 6 . 0 1.165 1.165

101

TABLE 28

NORMALIZED ABSORBANCE DATA FOR CoBr3" AT 704 nm

"(Normalized to C„ ' = 0.0030F)

CCo °Br A(1.0 cm) A(10 cm) A.......n

2 . 0 x 1 0 ” 3 6 . 0 x 1 0 “ 3 1..314 1.970

2 . 2 6 . 0 1 . 1 1 0 1.513

2.3 6 . 0 1.078 1.405

2.4 6 . 0 1.037 . . . . • 1.297

2 . 6 6 . 0 0 . 8 6 8 1.005

2.7 6 . 0 0.855 0.950

3.0 6 . 0 0.775 0.775

2 . 0 x 1 0 ” 4 6 . 0 x 1 0 " 4 1.274 1.909

2 . 2 6 . 0 ’ 1.105 1.507

2.3 6 . 0 1.078 1.405

2.4 6 . 0 1.030 1.288

2 . 6 6 . 0 0.800 0.922

2.7 6 . 0 0.787 0.875

3.0 6 . 0 0.653 0.653

The average value of [Br~] for several corresponding solutions

where n is close to 2.5 , is 13.9 x 10” at 680 nm and 9.0 x 10” at 704 nm

The calculated value of or l/[Br“] is 0.9 x 10"*.

Spectra of Cobalt Iodide Solutions

The spectra of cobalt Iodides In acetone, are shown in figure 21.

„ It will be shown that, of the halides, iodide alone appears to form a sig­

nificant concentration of monohalide complex. Within the solubility limits

of potassium iodide in acetone there appears to be no tetraiodide complex

formed. Absorptivities in the red continue to rise as the halide ion

becomes heavier, and the absorbance peaks move to higher wavelengths. The

iodides all show considerable absorption around 400 nm. This is probably

due to a charge transfer phenomenon, which is often more prominent with

metal iodides. The extreme case occurs with copper(ll) in aqueous solu­

tion. The electron does not jump back and forth between the cation and

anion, but stays on the cupric ion, and the copper(ll) is reduced to

copper(l).

It is possible to determine three stability constants, K^, 1^,

and for cobalt iodides. The first two overlap, but not as completely

as the corresponding chloride and bromide constants. All three constants

can be determined by the method of corresponding solutions. One example

of corresponding solutions from actual experimental absorbance curves has

already been presented in figure 16. Curves for normalized absorbance

data at two iodide concentrations are shown in figure 22.' At 735 nm or

any other wavelength above 700 nm, two definite changes of slope can be

detected. These occur where the iodide to cobalt ratios, are approximately

1.2 and 2.3. These slope changes just beyond integral values indicate

that three complexes, CoI+ , Colg, and Col^” , are formed. The slope

decreases after the ratio reaches 3.0 as the formation of Col^"" is

gradually completed.

FIGURE 21

ABSORBANCE OF COBALT(ll) IODIDES IN ACETONE

Ratio of Iodide to Cobalt as Indicated

1.80

3.61.40

\2 .7

1.00

.1,750.60

0.20

760 640 nm720 680

CORRESPONDING SOLUTIONS FOR Co-I SYSTEM IN ACETONE AT 735 nm

1.4

0.6

0.2

2,00.4

-4

FIGURE 22

Absorbance data for the cobalt iodide system in acetone are

presented at five different wavelengths in tables 29 to 33. For

calculations of K^, and Kg one must work on either side of the broad

absorbance plateau which occurs from about 665 to 705 nm. Calculations

show that [l~] is 3.4 x 10”^ at ii = 0.5 and 1.9 x 10”^ at n c 1.5. The4reciprocals for these concentrations give 3.0 x 10 as the preliminary

4value of K^ and 5.2 x 10 for Kg. The fact that the preliminary Kg is

greater than K^ indicates considerable overlap in the equilibria.

Further calculations by a more refined method are necessary to calculate

the exact values of K^ and Kg. The meaningful constant, ^2 , which can be9compared with the other halide constants, is K^Kg or 1.5 x 10 . Alter-

2 - 'nately, f3g can be calculated as l/[l ] where ii = 1.0. This concentra--5 9tion is 2.6 x 10 , so again (3g is calculated to.be 1.5 x 10 •

In order to calculate K^, one may use the I:Co ratios between

2.5 and 3.6. Even though the curves flatten above a ratio of 3.0, the

horizontal tie lines between normalized absorbances still indicate two

corresponding solutions. The average of many calculations yields a

yalue of [I ] = 5.4 x 10 ^ at n = 2.50, yielding K^ = 1.9 x 10^. Thus,

the values of the iodide stability constants are a little lower than

those of the corresponding bromide constants. No appreciable concentra­

tions of the tetra-iodide form in a saturated solution (0.075M) of

potassium iodide in acetone. Lithium iodide was not used because it is

not available in anhydrous form, and it cannot be dried without decomposition.

A summary of the stability constants obtained in this laboratory

and by Fine (1962) appears in table 34.

106

TABLE 29

' ABSORBANCE DATA FOR' Co-I SYSTEM AT 664 nm(Normalized to C *= 0.0018F) t»o .

. CCo CI. . A(1.0 cm) A(10 cm) An

7.2 x 10" 3 3.6 x 10" 3 0.645 0.1614.8 3.6 0.657 0.2463.6 3.6 0.670 0.3353.0 3.6 0.660 0.3962.4 3.6 0.655 0.4912 . 0 3.6 0.640 0.5761 . 8 3.6 0.650 0.650

. 1 . 6 3.6 0.636 0.7281.44 3.6 0.590 0.7391.32 3.6 0.552 0.7531 . 2 0 3.6 0.495 0.7421.08 3.6 0.447 0.7451 . 0 0 3.6 0.407 • • • • • 0.7327.2 x 10“ 4 3.6 x 10” 4 0.640 0.1604.8 3.6 0.640 0.2403.6 3.6 0.658 0.329

* 3.0 3.6 0.660 0.3962.4 3.6 0.682 0.5122 . 0 3.6 0.665 0.5891 . 8 3.6 0.650 0.6501 . 6 3.6 0.640 ' 0.7201.44 3.6 0.595 0.7451.32 3.6 0.548 0.7471 . 2 0 3.6 0.448 0.7231.08 3.6 0.434 0.7231 . 0 0 3.6 0.407 0.732

107

TABLE 30

ABSORBANCE DATA FOR CO-I SYSTEM AT 710 ran(Normalized to C_ = 0.0018F)Oo

CCo CI A(1.0 cm) A(10 cm) A.n

7.2 x 10~ 3 3.6 x 10" 3 0.660 0.1654.8 3.6 0.660 0.2483.6 3.6 0.680 0.3403.0 3.6 0.670 . 0.4022.4 3.6 0.692 0.5192 . 0 3.6 0.697 0.6271 . 8 3.6 0.705 0.7051 . 6 3.6 0.704 0.7921.44 3.6 0.712 0.8911.32 3.6 0.697 0.9511 . 2 0 3.6 .0.695 1.0431.08 3.6 0.680 1.1331 . 0 0 3.6 0.620 m • • m m 1.1167.2 x 10**4 , 3.6 x 10 4 0.568 0.1424.8 3.6 0.558 0.209

, 3.6 3.6 0.580 0.2903.0 3.6 0.585 ' 0.351*2.4 3.6 0.643 0.4822 . 0 3.6 '0.620 0.5581 . 8 3.6- 0.642 0.6421 . 6 3.6 0.652 0.7341.44 3.6 0.660 0.8261.32 3.6 0.623 0.8501 . 2 0 3.6 0.603 0.9041.08 3.6 0.565 0.9421 . 0 0 3.6 0.540 .0.973

TABLE 31

ABSORBANCE DATA FOR Co-I SYSTEM AT 735 nm(Normalized to C = 0.0018F) .......... Co

°Co CI A(1.0 cm) A(10 cm) An

7.2 x 10" 3 3.6 x 10“ 3 0.487 0 . 1 2 2

4.8 3.6 0.493 0.1853.6 3.6 0.500 .0.2503.0 3.6 0.504 0.3022.4 3.6 0.560 0.4202 . 0 3.6 0.608 0.5471 . 8 3.6 0.613 0.6131 . 6 3.6 0.633 0.7221.44 3.6 0.728 0.9111.32 3.6 0.803 1.0951 . 2 0 3.6 0.918 1.3771.08 3.6 * 0.935 • 1.5581 . 0 0 3.6 0.898 • • • • * 1.6177.2 x 10" 4 3.6 x 10 4 0.270 0.0684.8 3.6 0.272 0 . 1 0 2

3.6 3.6 0.300 0.150' 3.0 3.6 0.303 . 0.182.

2.4 3.6 0.360 .0.2702 . 0 3.6 .0.395 0.3551 . 8 3.6 0.435 0.4351 . 6 .3.6 0.491 0.5521.44 3.6 0.580 0.7261.32 3.6 0.605 0.8261 . 2 0 3.6 0.656 0.9841.08 3.6 0.658 1.097.1 . 0 0 3.6 ■ .... 0.653 1.175

109

TABLE 32

ABSORBANCE DATA FOR Co-I SYSTEM AT 745 ran(Normalized to C„ = 0.0018F) , - uo

o

. CCo CI.. . . .A(1.0 cm) A(l0 cm) A.n

7.2 x 1 0 “ 3 3.6 x 1 0 " 3 0.466 0.1174.8 3.6 0.480 0.1803.6 3.6 0.485 0.2423.0 3.6 0.492 0.2952.4 3.6 0.558 0.4182 . 0 3.6 0.638 0.5741 . 8 3.6 0.620 0.6201 . 6 3.6 0.660 0.7421.44 3.6 0.770 0.9641.32 3.6 0.875 1.1921 . 2 0 3.6 1.025 1.5371.08 3.6 1.055 ; 1.7581 . 0 0 3.6 . 1.005 • • • • • 1.8097.2 x 10“ 4 3.6 x 10~ 4 0 . 2 1 0 0.0534.8 3.6 0.225 0.0843.6 3.6 0.253 0.123

' 3.0 3.6 0.259 0.1552.4 3.6 0.313 0.2352 . 0 3.6 0.370 0.3331 . 8 3.6 0.415 0.4151 . 6 3.6 0.462 '■ 0.5201.44 3.6 0.592 0.740

- 1.32 3.6 0.635 0.8601 . 2 0 3.6 0.708 1.0611.08 3.6 0.720 1 . 2 0 0

1 . 0 0 .3.6 • '••••• 0.712 1.280

110

TABLE 33

ABSORBANCE DATA FOR Co-I SYSTEM AT 755 nm

(Normalized to C = 0.0018F)oo

°Co J CI A(1.0 cm) A(10 cm) A.n

1.80 x 1 0 “ 3 3.6 x 10“ 3 0.517 0.517

1.60 3.6 0.523 0.5881.44 3.6 0.665 0.8311.32 3.6 0.725 0.989

1 . 2 0 3.6 0.875 1.312

1.08 3.6 0 . 8 6 8 1.447

1 . 0 0 3.6 0.898 1.617

1.80 x 1 0 " 4 3.6 x 10” 4 0.330 0.330

1.60 3.6 . 0.416 0.467

1.44 3.6 0.495 0.6191.32 3.6 5.515 0.7021 . 2 0 3.6 0.575 . 0.8631.08 3.6 0.587 0.9791 . 0 0 3.6 5.580 1.043

Ill

TABLE 34

COMPARISON OF HALIDE STABILITY CONSTANTS IN ACETONE

.WITH.THOSE OF FINE *

System

CMGO. k3 ■ k4 ‘

(a) (b) (a) (b) (a) Cb)

Co-Cl

Co-Br

Co-I

2 . 6 x 1 0 9

1 . 6 x 1 0 9

1.5 x 109

93 x 10

2 x 1 0 9 .

> 1 0 9

1.5 x 105

0.9 x 105

1.9 x 104

> io5

> 1 0 5

2 . 2 x 1 0 4

215

65

. • • •

540

42

. 16

NOTE: (a) This laboratory, T = 25° (b) Fine (1962), T = 26°

No attempt was made to hold the Ionic strength constant, so

the figures are not true equilibrium constants. In general, one more

significant figure has been obtained in this laboratory. The signifi­

cant differences occur in the value of for the chloride and the in­

ability to obtain a value for the iodide. At an iodide concentra­

tion of 0.075M, there is only a hint that a tetraiodo complex is being

formed. The spectrum was not enough different from that of Col^” to

allow the calculation of the constant, K^. Another important differ-

ence is that a stable intermediate Col complex was detected in this+ * laboratory, and perhaps even a CoBr complex. There was no evidence of

a CoCl+ complex as was also true In the aqueous system.

The absorbance peaks and absorptivities based on the cobalt

concentration are listed in table 35 for all the halide complexes In

acetone.

112

TABLE 35

SUMMARY OF ABSORBANCES OF COBALT HALIDE

COMPLEXES IN ACETONE

Complex Absorbance Peaks, nm Absorptivity, e'

CoCl2 676, 578 300, 137

CoCl3" 690, 591 455, 235C o C l " 2A 701, 670, 623 590, 540, 333

CoBr2 679 387

■CoBr^” 702, 635, 619 665, 325, 303 .

CoBr,4 724, 697, 665, 640 1 0 2 0 975 610, 298

CoI+ 680 (broad) 185

CoI2 743, 706 348,' 393

CoI3" 745 850

Conductance Experiments In Acetone

If a salt, such as mercuric chloride, is weakly ionized, its

electrical conductance is low. In modern terminology, a weakly ionized

salt is a complex, so low conductance suggests that a complex is present.

The first experiment was performed to determine the relative

acidity of four inorganic acids in acetone. The relative conductance was

measured as a function of the current at a constant A.C. voltage. The

conductance cell had fixed platinum electrodes 1 . 0 cm in diameter and

113

2.0 cm apart. Therefore, both current and conductance are a reciprocal

function of the resistance of the electrolyte or 1 /R.

In the acid experiments 0.05 ml of the concentrated acid was

added to 10 ml of pure acetone. Then the current was measured at 25°C.

Relative conductance was calculated based on 0.06F hydrochloric acid as

1.0. The data appear in table 36. Based on conductance in acetone,

perchloric acid is about twenty times as strong as the other acids, all

of which are essentially completely.ionized in aqueous solution. The

' relative acidities of these four acids (HCIO^ > > HC1 > HNO^) are

about the same as those found by Kolthoff and Willman (l934) in glacial

acetic acid.

TABLE 36

RELATIVE CONDUCTANCE OF ACIDS IN ACETONE

Acid Formality Relative Conductance

HC1 0.06 1 . 0

hno3 0.08 0.3*

h 2 so4 0.09 1.3

HC10.4 0.06 23

Perchloric is the only acid tested which causes acetone to turn

yellow upon standing. This reaction must be related to the fact that

perchloric is the only common acid which is highly ionized in acetone.

Therefore, it appears that the reaction to produce the yellow color is

catalyzed by hydrogen ions. The color is first noticeable after two

hours, so it is not due to the simple addition of a proton to the

carbonyl group. The reaction still proceeds under a nitrogen atmosphere,

so air oxidation is not the primary cause. The color slowly disappearsu

when a salt, such as sodium sulfate, of an acid which is weak in acetone

is added, so the reaction is reversible. The most likely explanation of

the color reaction is that the acid catalyzes a polymerization reaction.

As a practical note, it is important that acetone solutions of perchlo­

rates should not contain any free perchloric acid. A saturated solution

of magnesium perchlorate did not produce any color.

Another set of conductance experiments waB run with cobalt

salts and lithium chloride. The results are shown in table 37, where

the conductance of 0.004F CoClg is arbitrarily set at 1.0. One can

observe that cobalt chloride, cobalt nitrate, and lithium chloride are ‘

all comparatively weak electrolytes in acetone. However, they are

stronger than the corresponding acids. This is another confirmation that

nitrate as well as chloride forms a complex with cobalt(ll) in acetone.

The mixing of LiCl and CoCl^ in a 1:1 ratio increases the conductance

above that of either salt alone, so an ionic complex, Li+CoCl- , appears/ Jto form. Addition of more lithium chloride to give a *Cl:Co ratio of 4:1

does not change the conductance much. The conclusion is that the chiefA _species present are still Li CoCl^ and excess LiCl.

The conductance experiments tend to confirm that the complexes,

CoCl^ and CoCl^ , form almost quantitatively in acetone. The tetrachloro

complex does not form quantitatively.

115

TABLE 37

RELATIVE CONDUCTANCE OF SALTS IN ACETONE

Salt Relative Conductance

CoClg, 0.004F 1 . 0

HC1, 0.06F 1.79

Co(N03)2, 0.004F 1.5

Co(C104)2, 0.004F 22.7

LiCl, 0.004F 1 . 6 •LiCl, 0.04F 7.1

LiCl, 0.004F, + CoCl2, 0.004F 6.9LiCl, 0.007F + CoCl2, 0.0035F 6 . 1

Solutions of Cobalt Chloride In the Lower Alcohols

In acetone it was shown that cobalt(ll) formed three different

complexes with chloride. Experiments in ethanol also indicate the forma­

tion of three cobalt chloride complexes, but in the more polar solvent,

the stabilities were not nearly as great.

• In methanol, the organic solvent most like water, only one

cobalt chloride complex could be obtained and identified by means of

spectrophotometry. The absorbance curves in methanol, shown in figure

23, are much like those of cobalt chloride in water. As in water, anhy­

drous cobalt chloride dissolves to form a pink solution. It takes a ten

to one ratio of chloride to cobalt to produce a visible blue tint. To

develop the blue color completely requires a Cl:Co mole ratio of over

116

FIGURE 23

ABSORBANCE OF Co(ll) + LiCl IN METHANOL (CCq =s 0.003F; CC 1 molarity as indicated)

/

\/v \

I \\m

'A

0.60-

0.20

740 700 660 620 580 nm

117

1000:1. Absorbance data at the maximum wavelength of 692 nm are listed

in table 38. The stability constant, calculated by the Ramette (1963)

method is 0*27.In ethanol some blue complex is formed when anhydrous cobalt

chloride is dissolved. In terms of polarity and structure, ethanol is

intermediate between methanol and acetone, so one would expect an inter­

mediate stability for the cobalt chloride complexes in ethanol. A family

of curves, shown In figure 24, indicates that three cobalt chloride com­

plexes are formed, and their spectra are similar to those In acetone. At

a Cl:Co ratio of 2:1 the shoulder at 580 nm Is evident. The maximum at

603 nm, characteristic of CoCl^"", does not appear until the Cl:Co ratio

is at least 20:1. The tetrachloro complex shows twin absorbance peaks,

like those in acetone, which appear at 663 and 689 nm. Three constants,

P2» K3» and K4» can be calculated- from the spectra. The chloride needed

to form CoCl^ was taken into consideration In the calculation of 3 ^ r

the value of x in the Ramette equation is expressed by

A W C CoCCl - CC12), . g ■

CCoCCl3The calculation of ^ yields 3.6 x 10 , and this figure may be used to

show that conversion to C o C ^ is 977. complete In 0.10M chloride solution.

Thus no more dichloride can be formed at the 20:1 and 50:1 ratios, so the

maximum at 603 nm must be from a new complex, probably CoCl^”. The

figure for the calculation (Whiteker and Davidson 1953) of K3 is 27, and.

for is 2.7. Absorbance data necessary for these calculations are

shown in tables 39 to 41. For the calculation of free chloride, it is assumed that the previous complex is completely formed.

1X8

TABLE 38

ABSORBANCE DATA FOR CoClg IN METHANOL

C ' Co CC1 cc i 2 . A692 .......

0.004 . 0.2 0.04 0.0400.004 0.4 0.16 0.111

0.004 0.8 0.64 0.4200.004 1.2 1.44 1.000.004 1.6 2.56 1.570.002 1.0 1.00 0.3330.002 1.4 1.96 0.6550.002 1.6 2.56 0.790.002 1.8 3.24 0.870.002 2.0 4.0 0.990.002 3.0 9.0 1.220.002 4.0 16.0 1.26

FIGURE 24

119

ABSORBANCE OF CoCll) +' X1C1 IN ETHANOL

(C = 0.003F; ratio of chloride to cobalt as given) Co

1.40-

1000

//

/ \

/ /

1.0 0 -

0.60-

-020

740 700 660 620 580 nm

120

TABLE 39

ABSORBANCE DATA FOR 02 FOR CoCl2 IN ETHANOL AT 662 nm

CC1 c 2 -Cl CCo A662

0.0040 1 . 6 x 1 0 “ 5 0.004 0.1180.0056 3.14 0.004 0.233

0.0068 4.63 0.004 0.332

0.0080 6.4 0.004 0.427

0.0096 9.22 0.004 0.545

0 . 0 1 2 0 14.4 0.004 . 0.74

0.0140 19.6 0.004 0.81

TABLE 40

ABSORBANCE DATA FOR K3 FOR CoCl3" IN ETHANOL AT 603 nm

°C1 CCo A603t

0 . 2 0 0.192 . 0.004 • 0.99

0 . 1 0 0.092 0.004 0.90

0.08 0.072 0.004 0.79

0.04 0.032 0.004 0.69

0 . 0 2 0 . 0 1 2 0.004 0.56

0.014 0.006 0.004 0.51

t

TABLE 41

ABSORBANCE DATA FOR K. FOR C o C l " 2 IN ETHANOL AT 689 nm4 4

CC1 i—i f i_i CCo A689

2 . 0 0 2 . 0 0 0 . 0 0 2 1 . 2 1

1 . 0 0 1 . 0 0 0 . 0 0 2 0.98

0.40 0.394 0 . 0 0 2 0.77

0 . 2 0 0.194 0 . 0 0 2 0 . 6 8

0 . 1 0 0.094 0 . 0 0 2 0.60

0.04 0.034 0 . 0 0 2 0.44

ffummary ■

A list of the halide stability constants obtained in organic

solvents at 25°C is shown in table 42. New contributions are the

constants in methanol and ethanol. Improved values have been obtainedi

for stability of the cobalt halides in acetone. The methods used

included two slope-intercept equations and the method of corresponding- solutions.

In any of the solvents used, including water, the highest

cobalt chloride complex has practically the same absorbance spectrum.<■ _2Sometimes the highest complex is CoCl^ ; sometimes it is CoClgCsolvent)

In any case the bond between cobalt and chlorine must be the same. The

structure must also be very similar, presumably tetrahedral.

122

TABLE 42

SUMMARY OF STABILITY CONSTANTS OF COBALT HALIDES

IN ORGANIC SOLVENTS

System Solvent • h K3 K4

Co-Cl Acetone1 2 . 6 x 1 0 9 1.5 x 105 . 215

Co-Cl Ethanol 3.6 x 103 27 2.7

Co-Cl Methanol 0.27

Co-Br Acetone 1 . 6 x 1 0 9 0.9 x 105 65

Co-I Acetone 1.5 x 109 1.9 x 104

CHAPTER V . *

THE DETERMINATION OF HALIDES WITH COBALT(ll) IN ACETONE

Introduction

Fine (1962, p. 1139) In his spectrophotometric experiments with

the halides of cobalt(ll) in acetone noticed that there is a linear

relationship between absorbance and the halide to cobalt ratio. There

is almost quantitative conversion to the dihalide, and then the third

halide ion in turn adds nearly quantitatively. Stability constants

determined by Fine and in chapter four of this work indicate that there

is about 1.5% dissociation when the cobalt concentration Is around

0.005M, the most practical concentration for spectrophotometric

determinations.

The possibility of applying this absorbance to a quantitative

method' for the determination of halides apparently was not suggested by

Fine (1962). In this application it became important to determine Iftthe slight dissociation was enough to Interfere with a quantitative

determination under practical experimental conditions.• *

Early experiments in this laboratory indicated that within

experimental error absorbance in the red is a linear function of the

chloride content as long as the Cl:Co mole ratio is less than 2:1.V >Bromide ion in the presence of cobalt(ll) in acetone behaves in a similar

123

124

manner. The absorption peaks are 675 run for chloride and 677 nm for

bromide. These results indicate that either chloride or bromide forms

CoX^ without the formation of any measurable intermediate, CoX+ .

The results with iodide are somewhat different. Relatively

small concentrations of iodide give a broad peak at 680 nm, probably due

to Col . As the I:Co ratio approaches 2:1, the absorption peak shifts

to over 700 nm, probably due to Col^*

If the cobalt concentration is constant, one should be able to

measure the chloride or bromide concentration by this new method.

Apparently the sensitivity is superior to that of most other spectro­

photometric and volumetric methods for halides.

The classical reagent, silver nitrate, has been adapted for the

determination of chloride in many ranges and matrices. Macro amounts may

be determined either gravimetrically as AgCl or by titration with

standard AgNO^. New methods and reagents offer little improvement,

except that potentiometric titrations with a recorder save operator time

in routine analysis. Complexometric titrations with mercuric nitrate to

form HgCl^ may be superior because there is no precipitate to obscure a

visual end point (Dubsky and Trtilek 1933, Cheng 1959)'.

Quick routine approximate determinations of chloride have been

made by determining pCl with a AgCl-Ag electrode, by determining the

potential* versus that of a calomel reference electrode (Malmstadt, Fett,

and Winefordner 1956; Stern et al. 1958), using an ordinary pH meter.

Recently Van Loon (1968) has developed a solid-state membrane-type AgCl

electrode for rapid determinations of chloride (10”^ to 10”^M). Bromide .

and iodide do not interfere if first oxidized by chromic acid. An AgBr-Ag

125

_5electrode has been used for measuring as little, as 10 M bromide (Pflaum,

Frohliger, and Berge 1962),Cathodic stripping has also been applied. Chloride is deposited

oas Hg^Cl^ on a mercury (Ball, Manning, and Menis 1960) or as AgCl on a

silver (Laitinen and Lin 1963) anode. In either case the polarity is

then reserved, and the metal chloride contributes to the cathodic current

as the metal ion from the precipitate is reduced.Some indirect color methods are based on the use of silver ion

as a reagent. One example is the spectrophotometric determination of the

chromate ion (Boltz 1958) released in the reaction,-2Ag2CrO^ + 2C1 -♦ 2AgCl + CrO^ .

An even more indirect method involves the reaction of chloride with silver

phosphate (Boltz 1958). The released phosphate ion is determined as the

molybdenum blue complex.

The silver chloride nephelometric method is sensitive to micro­

gram amounts of chloride (Boltz 1958). The amount of absorption is not

reproducible, for it varies widely with ionic strength and temperature.

i There are also several indirect color methods based on the use

of mercury(II) as a reagent. The mercury(II) ions react with chloride

ions to form soluble but slightly dissociated HgC^. Excess mercuric

ions can be determined colorimetrically as the diphenylcarbazone complex (Gerlach and Frazier 1958), where the sensitivity is great (e = 19,000),

but the complex decomposes after two minutes. Measurement of the color

vof chloranilic acid, freed from mercuric chloranilate by chloride (Barney

and Bertolocini 1957), is another method. Sensitivity is said to be about one part per million.

Many analysts have used the mercuric thiocyanate-ferric ion

reaction (Iwasaki, Utsumi, and Ozawa 1952; Elsheimer, Johnston, and

Kochen 1966; Rowe 1965;.Bergman and Sanik 1957). Chloride ion releases

thiocyanate ion from HgCSCN^ to form HgCl^. The thiocyanate ion in

turn forms the familiar red complex, Fe(SCN)++, with ferric ion.

A quite different approach involves gas chromatography

(Bergman and Martin 1962). Acid gases are distilled from 80% sulfuric

acid, and they reach the detector in the order: CO,,* ®2^* an<* HBr.One of the most sensitive methods of all is neutron activation

analysis. By this method Cosgrove and coworkers (1958) could detect-3iodide and chloride down to one microgram and bromide to 1 0 microgram.

Bromide can be concentrated even further by precipitating it as silver

bromide, and then irradiating it with neutrons (Filby 1964; Ballaux, Dams,81and Hoste 1967). Bromine has a high nuclear cross section, 3.1 barns,

82forming Br , which has a fairly long half life of 35.9 hours. By an ( ‘

even more sophisticated method, Hull and Gilmore (1964) used nuclear

activation and then an IBM 7094 computer to sort out the peaks' of chlorine

jind other elements in lubricating oils. A big advantage of neutron

activation is that usually no prelimianry separation of the-constituent

to be determined is needed.

References for the direct analysis of halides by measuring the

color of metal halide complexes are few. One is the measurement of an

FeCl++ complex (West and Coll 1956) in perchloric acid at 350 nm. The

.measurement must be carried out in 8 F perchloric acid, and the sensitivity

is not very great. Halide complexes are also formed with Pd(ll), where

the absorbance of the monochloride or bromide complex is made at 230 nm

'*■ ' ..." 127

(Chapman and Sherwood 1957; Weed 1964). Again measurements are only

moderately sensitive, and many other Ions, which absorb in the ultra­

violet, interfere.O '

Bromide and iodide may be oxidized to the free elements,

extracted with carbon tetrachloride, and the color measured in the

organic layer. Iodide may first be oxidized with nitrite, and then

bromide with potassium permanganate (Filby 1964).By a novel method Rudolph and Nadalin (1964) concentrated

chloride from titanium sponge as AgCl and then determined it by X-ray

fluorescence of the silver. Garska (1968) modified the method for

determining raicrogram amounts of chloride in silver catalysts.

Another reagent which may be adapted to indirect volumetric

analysis of halides is mercuric oxycyanide (Vieb’ock 1932; Belcher,MacDonald, and Nutten 1954). Halides react as follows:

*

HgO.Hg(CN) 2 + 2X" + HgO - HgX2 .Hg(CN) 2 + 20H-.

The liberated base is then titrated with HgSO^.

In all of these studies no one has mentioned using the forma­

tion of a colored halide complex in an organic solvent, although organic /

solvents were sometimes added to depress the ionization of mercury and

silver halides. Therefore, the use of a cobalt halide complex in acer-

tone appears to be the basis of a new analytical method.

In the determination of halide in an actual aqueous sample,

evaporating away the water presents no problem in dilute solutions of •

alkali halides. The dry halides are extracted with an acetone solution

of cobalt perchlorate, and for chloride the absorption of the blue CoCl2

complex is measured at 675 nm.

• . }

128

In addition to aqueous solutions of halides, one should be able

to decompose organic substances containing halides. Most organic halides

can be converted to alkali saltslu so that the final determination is

essentially inorganic.Fortunately, there have also been advances in methods for the

decomposition of organic compounds for analysis. Fusion with sodium

peroxide in a Parr bcmb results in complex decomposition, but extraction

of traces of sodium chloride with acetone from the large mass of other

salts seems impossible. The reduction with sodium or potassium metal

Btill adds too many alkali salts. However, two methods of destruction

add nothing but oxygen and water. One involves the burning of the organic

compound in a tube furnace in a manner similar to the combustion method

for carbon and hydrogen (Belcher and Ingram 1952; Haspanti 1967). The

disadvantage of .this method is that the halides are oxidized mostly to

the free halogens. In the Schoniger method the combustion with oxygen

is performed in a closed flask (Sch'oniger 1955 and 1956).

With the newer adaptation of electrical firing, the Schoniger

bombustion is easy to carry out. The small amounts of chlorine formed

can be reduced with (Schoniger 1955). • Bromine may be reduced with

nitrite (Awad et al. 1966) or bisulfite (Belcher and Fildes 1961).

Thus chlorine and bromine in most organic compounds can be

converted to the simple halide ions. Iodine is oxidized mostly to

iodate, which is more conveniently determined by an iodometric titra­

ction. Carbon and hydrogen burn to harmless carbon dioxide and water.

Sulfur burns mostly to sulfur dioxide, which is helpful in that it reduces

halogens to halides. Phosphorus burns mostly to P4°10 and finally forms

129

phosphoric acid which does not interfere. The behavior of nitrogen

--depends upon how it is combined in the organic compound (Fildes and

MacDonald 1961). Amines are converted to harmless nitrogen gas. Nitro4

groups become NC^ gas, which is converted to a mixture of nitrites andO*nitrates in water. Most heavy metals form oxides, which are insoluble

in dilute alkali, the absorbent for the combustion gases. These remain

in the flask while the absorbent is decanted away. .

Of the above substances likely to be present after the combus­

tion of an organic compound, only nitrates and zinc were shown to inter­

fere with the halide determination. Though not tested, cadmium and mer­

cury may be presumed to interfere also due to complex chloride formation.

Fluoride does not form a complex with cobalt in acetone and is harmless.

There would be mutual interference if more than one halide were present.

Alkali must be added to the absorbent to keep the halides in

solution during the evaporation step. Lithium hydroxide is preferred

because its chloride is the most soluble in acetone.

The cobalt chloride in acetone method appears to be practical

for two types of samples: aqueous solutions and most organic compounds

where only one halogen is present. Except for a spectrophotometer,

which is available in nearly every analytical laboratory, no expensive

or complicated equipment is needed.

Experimental

Apparatus: Spectrophotometer, preferably with an infrered photo-detector

for measurements at 675 nm. The Beckman DB Spectrophotometer was used. Schoniger combustion flasks with attachments for electrical firing.

, . 130

Reagents; Lithium hydroxide, 0.02N, in water

Cobalt perchlorate, 0.005F in acetone. This should not contain free

perchloric acid. u

Standard halide solutions, 0.01F in water; make from sodium or potassium

salts.

Reducing agent, 307., for chlorides; and NaHSO^, 0.01F, for bromides.

Pure oxygen gas.

Procedure: Take a sample containing from 5 to 50 micromoles (for chloride,

0.2 to 2.0 mg) of either chloride or bromide. Larger samples may be used

if the final dilution is proportionately greater. Water samples, if acid,

should be treated with alkali to prevent loss of halogen acids during the

evaporation step. Organic samples must be burned in such a way as to add

no large quantity of alkali salts, preferably by the Schoniger combustion

method. Liquids may be weighed in a gelatin capsule, but a blank must be

run on an empty capsule.

Weigh an organic sample and wrap it in filter paper. Place the

^aper and sample in the platinum basket which hangs from the stopper of

the flask. Make a rough calculation of all salts that will be present,

and then add to the flask 1.5 times the amount of lithium hydroxide

solution necessary to combine with all anions. The final volume should

be at least 10 ml. Replace the air in the flask with oxygen. Stopper

the flask, and fire according to the directions given for the equipment used.

After firing and cooling, transfer the flask's contents into a

small beaker. Rinse with dis.tilled water, and add the rinsings to the

131

~ beaker. To determine chlorides, add two drops of 30% hydrogen peroxide

as .a.jreducing agent for hypochlorites. To determine bromides, add an

equimolar quantity of sodium bisulfite. If the organic compound alreadyo

contains sulfur, no additional reducing agent is necessary.

Evaporate the solution to dryness at moderate heat, about 90°C.

Let the beaker cool, and add 5 ml of cobalt perchlorate in acetone.

Swirl the solution a few times, and allow to stand for ten minutes. Do

not try to break up the salts mechanically, for colloidal hi.^0^ will

.become -dispersed in the acetone and will interfere with the absorption

measurements. Transfer the acetone solution to a 10-ml volumetric flask.

Place another 4 ml of cobalt perchlorate solution in the beaker, and

allow another ten minutes for extraction of the halides. Transfer to the

same volumetric flask. Rinse the beaker with one milliliter of cobalt

perchlorate and transfer. Usually a few more drops of solution will be

necessary "to'rieach- the mark of the volumetric flask.

Transfer some of the solution to a one-centimeter spectro­

photometric cell. Measure the absorbance at 675 nm for chloride and at

-677-nm for bromide. Compare this absorbance with those obtained from

— standard halide solutions. Use distilled water in the reference cell.

Preparation of standard curves; From 0.01F solutions make a.series of

samples containing from 5 to 60 micromoles of halide. To each add 1.5

times the number of moles of lithium hydroxide. No reducing agent is

needed. Evaporate to dryness and carry through the procedure as out­

lined above. The standard curves are shown in figures 25 and 26.

132

FIGURE 25

DETERMINATION OF CHLORIDE IN ACETONE AT 675 ran (cCo = 0.004F)

0.9

0.7

-----A

0.5

0.3

2.01.60.8 1.2O 0.4C l'/ Co + +

• '133

FIGURE 26

DETERMINATION OF BROMIDE IN ACETONE AT 677 nm(CCo = 0,004F)

1.3

0.9

0.7

0.5

0.3

6.4 8.03.2 Br “

Cobalt perchlorate is now available from Alfa Inorganics, Inc.,

Beverly, Massachusetts. It can be made by adding a slight excess of

perchloric acid to cobalt carbonate. Cobalt perchlorate is easily soluble

in acetone and forms a clear pink solution. Presumably only the solvated

complex, Co(C3H 6 0 )g++, is present as the absorption curve is similar to

that of cobalt perchlorate in water.After a few days the acetone solution turns yellow, and there

is considerable absorption in the 400-A50 nm region. The solution can

still be used for chloride determination since there is no absorption

above 670 nm. The cause of the yellow color is a puzzle, however. The

color develops when no cobalt is present, but hydrogen ion appears to

be a catalyst for the reaction, which was discussed in more detail in

chapter four.As a practical note, it. is best to evaporate away as much of

the excess perchloric acid as possible. Yet one must stop the evapora­

tion before black oxides, of cobalt begin to form.

Results and Discussion

Comparison of the bromide and chloride absorbance curves; The wavelength

of the absorbance peaks for cobalt bromide and chloride are nearly iden-

'tical, 677 nm for bromide and 675‘nm for chloride. The bromide, however,

has a greater molar absorptivity at its maximum; e = 387 for bromide, and

300 for chloride. Both curves are shown in figure 27. The chloride curve

differs in another respect. There is a minimum absorption at 592 nm and

a slight peak at 578 nm. Since the cobalt bromide' curve lacks this

second peak, this feature permits a qualitative distinction between

135

. FIGURE 27

ABSORBANCE OF CoXg IN ACETONE (CCo = 0.004F; Cx = 0.005F)

/////7

\ 740

700

660

620

580

540

■ . ■' _ • 136- "4

chloride and bromide. The curves hold to the previously described shape

only as long as the mole ratio of halide to cobalt is 2:1 or less. More

halide produces the complexes CoX^” and CciX^ , which are discussed in

chapter four.

The iodide absorbance curves: Iodide appears to form both a mono- and a

di-iodide complex with cobalt. The mono-iodide has a distinct peak at

665 nm and a broad peak between 690 and 702 nm. As the iodide to cobalt

ratio passes 1:1, the peak at 665 nm becomes a shoulder, but the plateau

around 700 nm remains. A new peak at 742 nm appears. The curves are

shown in figure 28, where the iodide concentration is held constant.

The absorbance curves of cobalt iodide form an isosbestic point

at 675 nm and this point can be used for approximate iodide results. A"~

family of curves for solutions in which the cobalt concentration was

kept constant is shown in figure 29. However, either iodine or iodate is

difficult to reduce quantitatively in neutral or alkaline solution, and

there are probably better ways for determining this element.

‘ The oxidation of iodide to iodine by iodate yields six equiv­

alents of iodine for every equivalent of iodide originally present

through the reaction,

103" + 51* + 6H+ -• 3I2 + 3H20.

The liberated iodine may be determined colorimetrically by the familiar

starch-iodide method. Larger quantities of iodine are titrated with

thiosulfate.

Interferences: Unfortunately, nitrates interfere. Nitrate ion also forms

an acetone-soluble complex with cobalt, and there Is competition between

137

FIGURE 28

ABSORBANCE OF COBALT IODIDE IN ACETONE AT CONSTANT IODIDE CONCENTRATION(C_ = 0.005F)" I

Ratio of Iodide to Cobalt as Given

2.Q0.90

0.70 0.5

0.50

0.30

0.10

760 720 680 640 nm

•138

FIGURE 29

ABSORBANCE OF COBALT IODIDE IN ACETONE AT CONSTANT COBALT CONCENTRATION

(C^o = 0.004F; Cj as Given)

1.0 0 -

0.007

0.004

// 0.002/

QP.03........... \.............

0.60-

Q20-

760 720 680 640 nm

nitrate and halide ions for cobalt. Some absorbance curves for cobalt

chloride in the presence of nitrate are shown in figure 30. Mixed

complexes result with absorption maxima anywhere between 570 and'630 run,

depending on the relative amounts of halide and ni'trate present. Other

evidence for the formation of a cobalt nitrate complex is that for the

pure salt in acetone, the absorption maximum is at 536 nm and 6 — 52.

For fresh cobalt perchlorate solution, the absorption maximum is at

518 nm and e = 12. Fortunately, most organic nitrogen becomes nitrogen

gas after combustion (Fildes and MacDonald 1961), and thus causes no

interference.

Other common inorganic anions do not interfere unless the

total salt concentration is too high. Some data are presented in table

43. Sulfate, sulfite, and phosphate do not interfere as long as the

molar ratio of oxy-anion to halide is 1:1 or less. On a weight basis

five milligrams of the sodium salt can be present. The molar ratio of

fluoride to other halide may be as much as 2:1. The alkali metals,

alkaline earths, and ammonium ion do not interfere. Most heavy metals

precipitate as oxides during the alkaline evaporation, and thus are*

removed from further reaction. Zinc, cadmium, mercury, and- silver

interfere by forming their own complexes or precipitates with chloride.

Accuracy and precision: Varying concentrations of standard halide

solutions were taken as samples and analyzed according to the suggested

procedure. The results are shown in tables 44 and 45. The accuracy

and precision are as good as those for other known methods. For instance

a relative standard deviation of 2 .8% at the 1 . 0 mg level for chloride is

140

FIGURE 30

EFFECT OF NITRATE ION ON THE ABSORBANCE OF COBALT CHLORIDE (Co++ = 0.008F; Cl" = 0.010F; N03" as Given)

•-N

Ec

O oo

oCDo

oOJa

740

700

660

620

580

540

141

TABLE 43

EFFECT OF OTHER ANIONS ON THE DETERMINATION OF HALIDES

Halide Added Other Salt Amount Added, mM Absorbance % Recovered

Cl, 0.04 mM None • • • • 0.525 • • « •

i t KN03 0 . 1 0 0.370 70.5i t II 0.04 0.412 78.5i t MgS04 0.04 0.525 1 0 0

i t (nh4 )2hpo4 0.04 0.530 1 0 1

it NaF 0.08 0.515 98i t NaHS03 0.04 0.505 96

Br, 0.04 mM None • • • • 0.655 • • • •

ti kno3 0 . 1 0 0.258 39.4i t It 0.04 0.432 6 6

i t MgS04 0.04 0.670 1 0 2

n (nh4 )2hpo4 0.04 0.635* 97i t NaF 0.08 0.675 . 103i t NaHS03 0.04 0.650 99

142

TABLE 44

SUMMARY OF CHLORIDE RESULTS

Cl Taken, mg Cl Found, mg Error, mg

0.177 0.185 +0.008

0.355 0.362 +0.007

0.532 0.518 -0.014

0.709 0.706 -0.003

1.063 1.063 0 . 0

1.419 1.450 +0.031

1.773 1.832 +0.059

2.13 2.23 +0 . 1 0

NOTE: If last result is discarded, the standard deviation is 0.028 mg.

X43

TABLE 45

SUMMARY OF BROMIDE RESULTS

Br Taken, mgi

Br Found, mg Error, mg

0.400 0.336 -0.064

0.799 0.767 -0.032

1 . 1 0 0 1.174 -0.025

1.559 1.559 ‘ 0 . 0

2.40 2.49 +0.09

3.20* 3.15 -0.05

4.00 4.02 +0 . 0 2

4.79 4.81 +0 . 0 2

NOTE:. The standard deviation is 0.057 mg.

4

as good as one can expect for a spectrophotometrie method. The precision

on the weight basis for bromide appears to be poorer simply due to the

higher atomic weight of.bromine.

Comparison with volumetric method: A sensitive visual volumetric method

was compared with the cobalt chloride spectrophotometrie method. The

titration was made with silver nitrate by Fajan’s method, which uses

dichlorfluorescein indicator in neutral solution. In order to increase

sensitivity, by reducing the solubility of silver chloride, the chloride

was dissolved in 907. acetone. The results are compared in table 46. The

new method is preferable in every respect except time.

TABLE 46

COMPARISON OF CoCl2 METHOD WITH AgNO^ TITRATION

* CoCl2 AgN03

Blank none 0.25 mg

Practical minimum 0.18 mg 0.355 mg

Standard deviation 0.028 mg 0.037 mg

Time after sample preparation 30 min. 30 min.

Adherence to Beer's Law: The working curves for chloride and bromide are

shown in figures 25 and 26. The absorption curve for chloride at 675 nm

deviates only very slightly upward from a straight line below a Cl:Co

ratio of about 1:5/1. Then there is a more marked increase in slope,

apparently due to the formation of some CoCl^*". The curve for bromide is

145

very close to a straight line below a Br:Co ratio of 1.5:1* Above this

ratio there is a decrease in the slope. The CoBr^ , which is presumed

to form, absorbs more strongly, but the absorption shifts to longer

wavelengths.

Additional details about the more highly halogenated complexes of3

cobalt were discussed in chapter four.

Summary

A new spectrophotometric method has been developed for the deter­

mination of chloride and bromide based on the absorbance of the CoCl^

complex in acetone. ' With a one-centimeter cell, the sensitivity to

chloride is about 1 0 parts per million on a weight basis in the acetone

solution. This compares favorably with the most sensitive volumetric

methods, and is more sensitive than any other known direct color or

spectrophometrie method. If the original sample is a water solution,

there is no limit to the volume that one can evaporate to dryness. The

The nephelometric silver chloride method is more sensitive, but its /poor reproducibility limits its practical use to qualitative or semi- quantitative work.

CHAPTER VI

u 'THE COMPLEXES OP TRANSITION METALS WITH POLYPHOSPHATES

Introductlon

The polyphosphates of transition metal complexes have been

Incompletely explored. This is especially true of the tetraphosphates

because of the difficulty of obtaining pure crystalline tetraphosphate

salts.

Many investigators have worked with pyrophosphates and tri­

phosphates, which are available commercially. Considerable interest

in polyphosphates has been spurred by the fact that adenosine di- and

tri-phosphates, ADF and ATP, and even the tetraphosphate are important

in biological oxidation and reduction reactions.' Metal complexes with

ADP and ATP often function as catalysts or as part of a larger enzyme

molecule. Miller and Westheimer (1966) studied the hydrolysis of Y-

^phenylpropyl di- and tri-phosphates, and found the reactions to be

similar to those, of ADP and ATP. In their first paper they studied

acid and base catalysis, and in the second they used an enzyme. They

observed that the hydrolysis of ATP mixed with luciferin produces light, while Y-ph®nlypropyl triphosphate does not.

Since an important part of the work in the study in this

laboratory concerns divalent-metal tetraphosphates, one of the problems

was the preparation of a pure crystalline tetraphosphate. Thilo and

146

. .. - • . 147

Ratz (1949) were able to prepare sodium tetraphosphate by alkaline

hydrolysis of sodium tetrametaphosphate, which is commercially available.

Their hydrolysis took 100 hours at 40°C. Unfortunately they could not

crystallize sodium tetraphosphate from solution.

Westman and Scott (1951) shortened the hydrolysis time by

raising the temperature to 70°C, but lower polyphosphates were also

formed due to further hydrolysis. They separated and identified the

various phosphate hydrolysis products by means of paper chromatography.

Quimby (1954) lowered the temperature for hydrolysis to 25-28°C

and used a 1007. excess of sodium hydroxide. The reaction took three

weeks, but stopped at the tetraphosphate.Pa012“A + 20H" - PA0 1 3 ~ 6 + H20

Quimby first prepared a crystalline tetraphosphate, namely

guanidinium tetraphosphate monohydrate, by precipitation from aqueous

formamide solution.

Several researchers (Griffith 1964, Schulz 1956, and Osterheld

and Langguth 1955) have been able to prepare solid lead tetraphosphate,

irhich is insoluble in water. For example, Griffith (1964) mixed lead

carbonate and 85% phosphoric acid, and heated the mixture to 550°C.

Then he prepared ammonium tetraphosphate by mixing the lead tetraphosphate

with ammonium sulfide. The insoluble lead sulfide was filtered off.

Thin (HH4 )6P4 0 1 3 -6H 20 was precipitated by adding an equal volume of

methanol.* The crystals of the ammonium salt were stable at room tempera­

ture for a year, but they decomposed at 50°C to lower polyphosphates.

Schultz (1956) also prepared solid bismuth tetraphosphate by heating

148

together bismuth oxide and phosphoric acid In a molar ratio of 1:4 at a

final temperature of 700°C.

Matsumoto (Watters and Matsumoto 1964) and Machen (1967) made

minor modifications in Quimby*s method (1954) of preparation of

guanidinium tetraphosphate. Machen (1967) also converted the guanidlnium

salt to the tetramethy1-ammonium salt by means of ion exchange. The

latter was stable for several weeks at 5°C as a concentrated solution.

Its use avoided the formation of guanidinium complexes with tetraphos­

phate.

Watters and his coworkers determined the acidity constants of

tetraphosphoric acid (Watters, Sturrock, and Simonaitis 1963). These,

especially and are necessary if one wishes to determine the

stability constants with metals by means of the pH lowering. Tetraphos­

phate, like other chain phosphates, forms complexes even with the alkali

metal ions. Matsumoto determined the stability constants of tetraphos­

phate with guanidinium (Watters and Matsumoto 1964) and with lithium,

sodium, and potassium (Watters and Matsumoto 1967).t If one can determine the free metal ion concentration in the

presence of a complexing ligand, he has a valuable tool for determining

the complexity constant. This was done by Machen (1967) by means of a

calcium-ion sensitive electrode. The polarograph was used by Watters

and his coworkers to determine the stability of raercury(l) (Watters and

Simonaitis 1964) and various copper(ll) polyphosphates (Schupp, Sturrock,

and Watters 1963; Watters and Matsumoto 1966; Sturrock, Loughran, and

Watters 1962). This is one of the few examples of a mercurous complex,

which normally forms only ionic complexes. Covalently bound ligands

149

would probably break the relatively weak s-p hybrid bond between the two

mercury atoms- The known stability constants of the copper(il) poly­

phosphates can be used in the indirect determination of the stability of

other metal phosphates, which are not'reversibly reduced at the dropping

mercury electrode.

Polyphosphates with more than four phosphorus atoms can be

prepared, but hydrolysis is too rapid at room temperature to permit the

isolation of any one of them in a pure state. Griffith and Buxton (1965

and 1967) have prepared long-chain polyphosphoric acids simply by

heating 857. H^PO^ to 400°C. After neutralization, a hexametaphosphate

(Griffith and Buxton 1965), Na^Pg O ^ g ^ ^ O could be extracted and crystal­

lized. Linear polyphosphates (Griffith and Buxton 1967) up to the octa-

phosphate were separated and identified by means of chromatography. The

average chain length was five. Decomposition was observed to occur in

two ways. A terminal FOg group, or less frequently, a trimetaphosphate

group split away. The pyro-, tri-,’and tetra-phosphates are stable in

neutral or alkaline solution, but hydrolyze slowly in acid. Longer-

<chain polyphosphates show significant decomposition within an hour at

any pH.

Gill and Riaz (1969) studied the kinetics of the degradation

of long-chain polyphosphates. They started with pure sodium "long-

chain11 polyphosphate, known as Graham1s salt. The coiled structure,

determined by X-ray diffraction, contains repeating units of three PO^

tetrahedra. These can form tridentate bonds with a metal cation, which

catalyzes the removal of a tremetaphosphate ring from the middle of a

chain. The rate of this reaction and the splitting off of a terminal

• ’ 150

orthophosphate group, both increase with increasing temperature and

acidity.Shen, Stahlheber, and Dyroff (1969) prepared and characterized

crystalline ammonium polyphosphates with a chain length greater than 50.

The long-chain polyphosphates are quite insoluble in cold water, and

decompose more slowly in contact with water than Graham’s salt.

Wu and coworkers (1967) showed by direct calorimetry that

hydrolysis of pyrophosphate to orthophosphate is an exothermic process.

They calculated the AH at a pH of 5 to be -4.5 kcal/mole at 25°C. Here-2the reaction was ^ O ^ P ^ (aq) + ^ 2° ^ 2 ^ 4 * 'r e react*on was cataylzed

by a phosphatase from Escherichia coli.

Generally the bonding of a metal ion is assumed to be made with

the oxygen from adjoining PO^ groups. One or more six-membered rings can

be formed. Since most transition metal ions are six-coordinate, the

remaining bond positions can be occupied by groups such as water or

hydroxide. The charge depends upon the metal \/,0TP,

ion, the polyphosphate, and the protonation of (H„0), M f0* 0-P

'end oxygens bound to phosphorus. A typical ^

complex is copper(II) monohydrogen pyrophosphate, CuHP^O^”.'Brintzinger and'Plane (1967) presented Raman spectra evidence

that pyrophosphate forms tridentate bonds with zinc and copper(ll) ions.

This means that one of the PO^ groups must furnish two bonds, just as

the NO^"* group sometimes does (Cotton, Goodgame, and Soderburg (1963).

Pyrophosphate appeared to be bidentate with Ni(ll), Co(ll), and Mg(ll).

Since there was no evidence of a metal-oxygen stretching frequency, the

M-0 bond was assumed to be ionic.

151

Experimental

The general method of obtaining data to calculate the stability

of a metal polyphosphate complex was to form the complex, and then

titrate it with standard acid. This titration curve was compared with

the one obtained when pure tetramethylammonium polyphosphate was titrated.

The degree of pH lowering during the titration was a measure of the

stability of the metal polyphosphate complex. These data were subjected

to a mathematical treatment, in which a slope-intercept equation was

developed for determining the two most important stability constants. To

eliminate activity coefficient effects, the ionic strength was kept

constant at 1 . 0 with tetramethylammonium chloride or nitrate.

Complexes of three linear polyphosphates were studied. Two,

the pyrophosphate and triphosphate, are available commercially as the

sodium salts. The third, tetraphosphate, had to be prepared in the

laboratory as the guanidinium salt.

The starting material for preparing guanidinium tetraphosphate,

was sodium tetrametaphosphate, which is available,

^from Victor Chemical Works as "Cyclophos." This was hydrolyzed and

converted to the guanidinium salt by the method described by Quimby (1954)

and Machen (1967). The crystalline guanidinium salt is stable indefinite­

ly at room temperature.

Machen*s method (1967) for the conversion of crystalline poly­

phosphate to the tetramethylammonium salt was modified somewhat. Two

millimoles of the sodium or guanidinium salt were carefully weighed and

dissolved in 20 ml of water. This solution was run through a 5-cm cation

. . 152

exchange column, Dowex 50W-X8, and the polyphosphate was converted to

the tetramethylammonium salt. The polyphosphate was eluted from the

column with 25 ml of 2M tetramethylammonium chloride. Enough tetra­

methylammonium hydroxide was added to bring the pH to 10.0 for the

pyrophosphate and triphosphate, and to 9.0 for the tetraphosphate. The

solution was finally diluted to 50 ml, and the ionic strength was adjusted

to 1 .0 .The next step was to add an equal number of moles of a metal

chloride or nitrate, Iron(ll) was added as solid FeCl^'^H^O. All other

metals were added as 0.10M solutions. Mixing caused the 1:1 metal-

polyphosphate complex to form, and the drop in pH was a qualitative

measure of the stability of the complex because the metal ion displaces

hydrogen ion in forming the complex.

The metal salt solutions were standardized as follows:

Magnesium, manganese, zinc, and lead were titrated with EDTA in a

buffered solution at pH 10 using the indicator, Eriochromeblack T. Iron

and aluminum were preciptitated with ammonium hydroxide and ignited to

the trivalent oxide. Nickel, cobalt, and copper were electroplated

on a platinum cathode, and weighed as the pure metal. Nickel was also

checked gravimetrically with dime thy lglyoxime..

After the metal polyphosphate was formed, it was titrated

with standard acid, usually 0.1N HC1 or HNO^. (The acid was standard­

ized against pure sodium carbonate.) At frequent intervals the pH was

recorded. The instrument used was a Beckman Research pH Meter, on

which one can read three decimal places. A Beckman glass electrode,

no. 40498, and a saturated calomel reference electrode were used.

. ' . 153

Titration curves, shown In figures 31, 32, and 33, were

obtained for some of the metal polyphosphates and the tetramethyl­

ammonium polyphosphates. One may observe that there is only a slight

difference in pH after two equivalents of acid are added per mole of

polyphosphate. The ratio between the numbers of equivalents of acid

added and moles of polyphosphate is defined as "a." The pH in the

solution containing copper is slightly displaced because the copper-3nitrate solution itself contained a little hydrogen ion (2 . 2 x 1 0

moles per liter).

The final volume after the titration was about 56 ml. This

included 50 ml of the original polyphosphate solution, 2.0 ml of the

metal salt solution, and 4.0 ml of the standard acid for titrating until

a = 2. A median total volume, or 53 ml, was used in the calculations.

Since the increase in volume during the titration was less than 87., the

effect of this change, which was largely cancelled, was considered to

be insignificant for most calculations.

The temperature during the titration was 25.0 ± 0.5°C, and was

^maintained by means of a water bath. All reagents were dissolved in .

doubly distilled and deionized water.

The possible formation of other complexes must be considered.

‘An example already mentioned is the formation of weak sodium polyphos­

phate complexes. These are avoided by conversion to the tetramethyl­

ammonium polyphosphate. The same conversion eleiminates guanidinium ion,

which,because of its similarity to ammonia, may complex transition

metals, and is known to form tetraphosphate complexes (Watters and

Matsumoto 1964). The chloride portion of this work showed that the

154

FIGURE 31

TITRATION OF M(ll) TETRAPHOSPHATES IN 1:1 RATIO

(CM « 0.00321M; C_ = 0.00322M; V = 53 ml)M L

10

8PH

MgL6

CuL4

2

31 2 4 5ml HCI, 0.1017N

155

FIGURE 32

TITRATION OF M(ll) TRIPHOSPHATES

(C„ = CT = 0.00377M; V = 53 ml)M L *

co

o CD C\JQ.

ml

HCI

0.10

72N

FIGURE 33

TITRATION OF M(ll) PYROPHOSPHATES

(CM = CL = 0.00240M; V = 52 ml)

156

C\J

O CD

CO

to

CO

c\]

Q.

ml

HCI,

0.05

08N

157

chloride complexes of cobalt(ll) and copper(II) are not stable in 1M

chloride solution. This is not true of ironClIl)^ and the stability of

zinc chloride was not tested. Therefore, the background electrolyteu

was changed to tetramethylammonium nitrate when iron(lll) and zinc

polyphosphates were titrated, and standard nitric acid was substituted

for hydrochloric acid.

Tetramethylammonium nitrate, which is not available commercial­

ly, was prepared as follows: four-hundred and sixteen grams of tetra-

methylammonium hydroxide, a 24% solution in methanol from Matheson,

Coleman, and Bell, were weighed. Slowly, boiled nitric acid was poured

in until the solution was acidic to methyl red. Then carbon dioxide and

most of the alcohol were removed by boiling. Next the pH was adjusted

to 7.0 with tetramethylammonium hydroxide, and the solution was diluted

to one liter. The final concentration was 1.1M.

In some cases, especially with pyrophosphates, there was

precipitation before or during the titration. This problem sometimes

could be avoided simply by diluting all solutions to one-half or one-

^fourth of the concentrations suggested earlier. The practical minimum

of metal ion and polyphosphate is about 0.001F. In any case, the ionic

strength was maintained at 1.0. Another precipitate was iron(lll)

hydroxide or hydroxypolyphosphate, but in all studies except with pyro­

phosphate, this precipitate dissolved after a short time. Owing to the

precipitation of pyrophosphates of cobalt, iron(ll) and (III), manganese,

and zinc, only approximate values could be obtained for the stability con­

stants of these pyrophosphate complexes. Lead ion formed a precipitate

even with triphosphate; further work with this element was not performed.

• 158

Notes on Experimental Technique

Nitrogen Atmosphere: Passing nitrogen gas through the solution removes

two undesirable gases, oxygen and carbon dioxide. Oxygen, of course,

must be removed when there is danger of air oxidation. In the presence

of oxygen iron(ll) is oxidized to iron(lll), and the difference in the

titration is several tenths of a pH unit. In order to eliminate oxygen,

nitrogen was bubbled through ferrous solutions for about ten minutes

before the titration began and then continued during the titration.

The other undesirable gas, carbon dioxide, reacts as an acid

and lowers the pH. In the case of nickel and cobalt polyphosphates, the

titration curves were the same whether nitrogen was used or not. There­

fore it was assumed that not enough CO,, was present to affect the titra­

tion. All solutions were made in water which had been initially freed

of carbon dioxide. -

Ion Exchange; The polyphosphates must be weighed as the sodium or guan­

idinium salts. Both of these cations form complexes with polyphosphates

,and may change the pH by as much as 0.5 unit. This is especially true, in

the case of magnesium, whose complexes are only a little stronger than

those of sodium.The problem was to remove the alkali metal and still have a

known concentration of polyphosphate in a small volume. A short cation

(Dowex 50W-X8) exchange column removed the sodium and let the polyphos­

phate. pass through. A concentrated (1 or 2F) solution of tetramethyl

ammonium chloride converted the polyphosphate to a form in which the

cation is noncomplexing. At the same time the total ionic strength was

159

adjusted to 1.0, Enough sodium ion remained to give the yellow flame

test, but no difference'in the titration curve was detected.

Effect of chloride: Another ion which complexes certain transition

metals is chloride. Yet because of its availability and greater

stability, it is more convenient to use tetramethylammonium chloride as

the background electrolyte. In the case of iron(lll) there is no choice.

One has to use a nitrate or perchlorate solution. The chloride portion

of this work has indicated that it should be safe to use 1M chloride in

the presence of cobalt(ll) but perhaps not for copper(ll). Data shown in

table 47 indicate that there is no significant difference between the

titrations in chloride and nitrate solutions. Perhaps the explanation

.is that there is so little copper ion uncomplexed by polyphosphate, that

the additional complexing by chloride affects the potential measurements

only insignificantly. The same titrating solution, standard nitric acid,

was used in all cases. Nickel was used as the reference metal because it

forms chloride complexes only in nonaqueous solvents.

Temperature: Since nearly all equilibrium data in this country are given/

at 25°C, this temperature was chosen for the titrations. In one case,

aluminum tetraphosphate, titrations were run at 24.5° and 27.5°c. Since

the two titrations curves were essentially identical, temperature effects

within a few degrees of 25°C appear to be negligible.

Points recorded during titration: After the addition of every 0.2 to 0.5

ml of standard acid, the pH was recorded. Later the points were plotted

on linear graph paper. Then the difference between the titration of the

160

TABLE 47

COMPARISON OF TITRATIONS IN CHLORIDE AND NITRATE SOLUTIONS

Phosphate BackgroundIon

pH at

a = 0.05 a = 0.75 a = 1 . 0 a = 1.5

Oo-P3°10 Cl- •6.005 5.624 5.263 4.422

■Co-P3°10 N0 3_ 6.009 5.609 5.261 4.421

Cu-P3°10 Cl- 5.641 5.159 4.640 3. 730

Cu-p3°10 N°3_ 5.652 5.136 4.628 3.739

N 1-P3°10 Cl- 5.993 5.640 5.310 4.537

Ni-P3°10 3 O U) 1 6 . 0 2 2 5.654 5.288 4.529

Co"P4°13 Cl- 5.9*57 5.633 5.324 4.640

Oo-p4°13 N0 3- 5.946 5.609 5.319 4.666

Nl-P4°13 C 13- 6.044 5.680 5.369 -4.668

, Ni-P4 0 1 3 N°3_ 6.064 5.693 5.362 4.664 .

• • 161

metal polyphosphate and of the polyphosphate Ion could be seen at a

glance. For'calculation one really needs only three points. These occur

at a = 0.5, 0.75, and 1.5. In any case the titration was stopped at

a « 2,5 or 3.

It may be noted that the titration of uncomplexed pyrophos­

phate and triphosphate gives two breaks. These occur at a = 1 and a s 2.

The.titration.of tetraphosphate gives only one break, where a = 2. In

any case all but the last two.hydrogens react like those of a fairly

strong acid. One weakly acid proton is located at each end of the poly­

phosphate chain. Metaphosphoric acids, which are cylic, have no weakly

acid proton. In fact, metaphosphoric acids are too strong for the metal

complex constants to be determined in the manner described here.

Order of addition of reagents: The first step is to prepare the

tetramethylammonium polyphosphate and adjust the pH. to 9 or 10 depending

upon the phosphate. The metal salt solution must be added slowly while

the polyphosphate solution is stirred. Otherwise a precipitate may form.

The solution should be clear before the titration is started.f ‘

Calculation of Metal-Polyphosphate Stability Constants

A variety of methods differing in mathematical detail are used

for determining metal-ligand stability constants from pH measurements

recorded during a titration. Several of these are summarized in a

recent publication from this laboratory (Watters and Machen 1968). For __

the cases in which an excess of metal ion results in precipitation it is

usually assumed that only the 1:1 complex is formed. On this basis

162

Ellison and Martell (1964) calculated the stability of several polyphos­

phate complexes. The method becomes less accurate when the consecutive

pK& values of the free acid are close together. This is especially true

for tetraphosphoric acid where pKflij and differ by only 1.7 units.

The method used here was modified from one used by Hammes and Morrell

(1964). Various mass-balance equations are developed and solved for

free ligand concentration. Finally a graphical solution is made for the

stability constants of the 1 : 1 complex of the nonprotonated and the

monoprotonated ligand. Concentrations of other metal complexes are

assumed to be insignificant. The method was used because, for the pres­

ent at least, more precise methods do not seem to be available for the

.studies of these complexes of many transition elements.

First, conditions are adjusted so that the only important

metal-ligand species are ML and MHL. This is generally true if the

total concentrations of metal ion, C^, and of ligand, C^, are approxi­

mately equal. If the pH is between 4 and 7.5, the uncomplexed ligand

exists as L, HL, and ^ L . (Charges are omitted for simplicity and

'because the free ligands have different charges.) In this pH range the

concnetrations of free H and OH ions are negligible.

The important stability constants may be expressed as follows:

[ML]Pioi “ ------- (29)1 0 1 [M][L]

[MHL]P1 U = --- :----- (30)Aii [M](H)[L]

[HL]Pnn ------ = 1/K„ (31)0 1 1 (H) [L] *

Expressions in brackets refer to concentrations in moles per

liter while the activities of H+ and OH are enclosed in parentheses.

The latter activities are measured directly by means of a pH meter.

Since two types of expressions are involved in these equilibrium constants,

they are known as ’'hybrid1' constants. The first number in the subscript

after P refers to the number of metal ions in the complex, the second to

the number of protons, and the third to the number of ligands.

The known quantities are the following:

C„ = added number of millimoles of metal ion/ml of solution. ’ M

CT = added number of millimoles of polyphosphate/ml of solution LC„ » added millimoles of HCl/ml of solutionII

* - V ° L

The various acid stability constants have been determined

independently (Watters, Sturrock, and Simonaitis 1963). For the

conservation of metal ion one writes

CM = [M] + [ML] + [MHL] . (33)

Substitute the equivalent values of [ML] and [MHL] from equations

(29) and (30),

CM = [M] + P1 0 1 [M][L] + PU 1 [M][L](H) (34)

Solve for [L],

- M[L] = ---------------------- (35)

PjOjM + Pm [M](H)

For the conservation of ligand one writes

CL = [M L ] + [M LH] + [ L ] + [H L ] + [H g L ] ( 3 6 )

CL " P l o i M m + P U 1 [M ](H )Q L ] + [ L ] + P0 U ( H ) [ L ] + { ^ ( H ) 2 [ L ] ( 3 7 )

Solve for [L],

[ L ] = .................................................. CL...................... ( 3 8 )

P 1 0 1 [M ] + P 1 1 1 [ M ] ( H ) [ L ] + [ L ] + Pq 1 1 (H)[L] + P0 2 l ( H ) 2 [ L ]

Equate expressions (7) and (10).

cm - M c l(39)

P lO iC M ] + P m [M ](H ) P 1 0 1 [M ] + P i n [M ] ( H ) + 1 + P X 11(H ) + PQ 2 1 ( H ) 2

Cross multiply and rearrange the terms. The result is a quadratic

equation in [M].

^ 1 0 1 + + Cl + Pq 11^H + ^ 0 2 1 ^ " CM ^ 1 1 1 ^ ** ^011CM +

^101C 1 + “ CM + CM ^ 0 1 1 ^ + CM ^ 0 2 1 ^2The solution for a quadratic equation of the form AM + BM + C = 0 is

DO - ~B ~ 4~ (4i)2A

Only the positive root is important in these calculations

165

Thus one can solve for [M] if he knows all the equilibrium

constants. The acid constants, (H) and C^, are known. The complex

formation constants may be estimated iteratively by using plausible

trial values of [M],

CH = [H+] - [OH”] + [HL] + 2[H2L] + [MHL] by definition (45)

In the pH range used, 4 to 7.5, the concentrations of hydrogen and

hydroxyl ions are insignificant, so

CH 8 [HL] +•2[H2L] + [MHL] = millimoles HCl/ml solution (46)

CH - Pq u ( H ) [ L ] + 2 P 0 2 1 < 1I)2 [ L ] + P1 U [M ]C H ) [L ] ( 4 7 )

Divide Cjj by C^. ■ •. ___________P0 1 1 <H> C ^ + 2 Pq2 1 ^H * Pu 1 [m](h)[l]

M ' i = 1' <5 1 ^ 0 /P i b i M M + P1 i;l[M](H)[L] + Pq1 1 (H)[L] + P0 2 1 (H) [L] + W

Cancel out [L], cross multiply, and rearrange the terms.

V M P l O l + (CH - Cl) M CH)PU 1 - (CL - CK)P0 U <H) + (2Cl - C ^ O O 2 Cg(49)

This equation is of the form, y =-mx + b, where

b « C^[M]Pioi = the y intercept

m = P j ^ = the slope

x = (CH - Cl )[M](H>

y = (CL - cH)Poil^H) + (2CL “ CH)P021^H ^ 2 ” CH

Thus one can solve graphically for P j ^ and P ] ^ hy solving for

x and y at two well-separated points on a pH titration curve. The best

points occur where a s 0.5 and 1.5, where a = equivalents of acid/moles

166

of ligand. For transition metals a plausible trial value of [M] is

around 57. of C^.The. slope is Ay/Ax, which gives P ^ directly. A more’

meaningful constant is

[MHL]

V = “ Plu/P°u (50)

The intercept, b, is calculated from the equation, y as mx + b.

In order to obtain more significant figures, the x and y values calculated

at a = 0.5 should be used. Then

^101 ~ KML =

These values of and P ^ are only approximate because a

trial value of [M] was used. The next step is to substitute these 0

values into the quadratic equation (41) for [M], One may note that [M]

is dependent upon (H) as well as on the stability constants. [M]

increases gradually as ’’a" increases from 0.5 to 1.2, then more steeply.

The best correlation with other workers is to select the (H) which

corresponds to a = 0.75. This is an empirical value, but one must /

remember that other complexes such as MgL and are form.ed» although

in smaller amounts, and these influence the titration curve.

With a better value of [M] one can recalculate x. Since the

y and b terms do not contain [M], they do not change, so it is easy to

calculate new values of P j ^ and P^q j - Continue recalculating [M] until

it does not change significantly. Actually the new 0 values, P^^^ or

^101* are simP1y t*ie old 0 values multiplied by old[M]/new[M],

Another note is that the initial volume is usually 50 ml, and

the volume of acid added at a » 1.0 is about 2 ml. Thus the volume of

acid added causes only an insignificant change in the concentrations of

and CL , and the effect of this slight dilution on both cancels.

Data for the complete titration of copper pyrophosphate are

listed in table 48.

TABLE 48

TITRATION OF COPPER PYROPHOSPHATE(n = CT = 0.00240F)Cu L

ml HC1, 0.0508N pH

0 . 0 6.8630.5 6.0781 . 0 5.6761 . 2 ' 5.5361.5 5.3381.7 5.2052 . 0 5,0132 . 2 4.8712.4 4.7422 . 6 4.6002 . 8 4.4643.0 4.3293.5 4.0153.7 3.9044.0 3.7464.5 3.4965.0 3.2865.5 3.117

NOTE: The solution became slightly cloudy after 3.0 ml of acid wereadded. The pH at this point was 4.33.

X' :

y = o

at a

y = -

168

Sample Calculation, Cu(ll) Pyrophosphate

Pyrophosphate, weighed as Na^P^O^*lOH^O, 0.0558^ = 0.125 iriM

V = 52 ml when a = 1.0; \1 ~ 1.0; T = 25.5°C

C_ = 0.125/52 = 0.00240JjCuCl^, added as 1.25 ml of 0.100 M solution, pH = 2.755

CM = 0.125/52 = 0.00240j,To change H activity to concentration, divide by 0.8

(H) = antilog(-2.755) = 1.76 x 10" 3

[H] = 1.76 x 10“3 /0.8 = 2.20 x 10' 3

H+ added with CuCl^ is 1.25 x 2.20 x 10” 3 = 2.75 x 10 3 mM

2.75 x 10~ 3 mM of H+ is equivalent to 0.05-ml of 0.0508N HC1

at a = 0.05 (1.18 ml) a = 1.5 (3.64 ml)

pH = 5.53 pH = 3.93

(H) = 2 . 9 5 x 10" 6 (H) = 1.122 x 10‘ 4

CH = Cl/2 = 0.00120 CH = 3/2 = 0.00360

PQ 1 1 = 108 -9 3 = 8.51 x 108

|3q 2 i = 101 5 -0 6 = 1.148 x 101 5 Let [M] = 1.0 x 10“ 4 at a = C^/C^ =* 0.5

= (CH - CL)[M](H) = - 0.0012(1.0 x 10-4 )(2.95 x 10“6) = 3.54 x 10" 1 3

y = (CL - Ch)Pq u (H) + C2Cl - Ch)P0 2 1 (H) 2 - Ch.0012(8.51x108)(2.95x10=6) + 0.0036(1.I48xl015)(2.95xl0'6)2 - 0.001

y = 3.02 + 35.9 - 0.001 = 38.9

= 1.5, x = 0.0012(1.0 x 10“4 )(1.122 x 10'4) = 1.348 x 10’ 1 1

0.0012(8.5lxl08 )(i.i22xl0"4) + 0.0012(1.I48xl015)(1.122xl0"4)2 - 0 . 0

y = 115 + 1.73 x 104 - 0.0 = 1.724 x 104

169

, 1.724 x 104 - 36slope = dy/dx =------ m — :-------itl1 1 1 1.348 X 10 lL - (-0.035 x 10 Ll)

P 1‘72Q x AP* . = 1.242 x 1 0 1 5 log P--, = 15.094. 1 1 1 1.383 x lO- 1 1

y = mx + b or b = y - mx

b = 38.9 - 1.242 x 101 5 (-3.54 x 10"13) = 38.9 + 440 = 479 = intercept

P 1 0 1 = b/ M CH " ------- V 1 ------- = 2 -0 0 x 1 q 9 1 o 8 P1 0 1 = 9.3011 0 1 H 1 x 10 (0.0024) 1 U 1 ■

With these trial values of P- q ^ an<l one may solve for amore accurate value of [M] by using the quadratic formula:

-B + V B2 - 4AC[M] =Let (H) = 1.0 x 10

2A -5

A ” ^111^ + ^101 A = 1.242 x 1015(1 x 10“5) + 2.00 x 109 = 14.42 x 109

B = 1 + eo u CH) + Po n (H) 2 + (CL - CM)CPl n (H) + Pl01)

B = 1 + 8.51 x 108(1 x 10“5) + 1.148 x 1015(1 x 10"5) + 0

B = 1.233 x 105

2 1 0 B b 1.522 x 10

C = -CM (l + P0 1 1 (H) + P0 Z1 (H)2} b -0.0024(1.233 x 105) = -296

r.,-, -1.233 x 105 + 1.522 x 101 0 -4(1.442 x 1010)(-296)W = --------------------------------rn--------------------

2(1.442 x 10iU)

-1.233 x 105- + 4.14 x 106 , • ,„-4[M] = ------------------ tv:------ = 1.393 x 102.884 x 10 U

170

Refine the value of x with the calculated [lQ.

x0 . 5 = -0.0012(1.393 x 10"4 )(2.95 x 10"6) = -4.94 x 10" 1 3

x x 5 - +0.0012(1.393 x 10“4 )(1.122 x 10“4) = 1.877 x 10- 1 1

1.720 x 104 1.720 X 104 0 ,rt14p = ------------ rr----------------- = --------------- rr = 8.93 x 10LLL 1.877 x 10 u - (-0.049 x lO-11) 1.926 x 10" 1 1

l°g P11]L = 14.951

The intercept, b, does not change.

p =------- ------------*Z|_--------c i.4 3 o x 109 log Pi m = 9.156[M]Cr 1.393 x 10 (0.0024)

Calculate the new 3 values by the shortcut method.

Pm = 1.242 x 101 5 x 1/1.393 « 8.91 x » 101 4

PlOi = 2.00 x 109 x 1/1.393 = 1.432 x 109

In further refining [M], only A in the quadratic formula

changes. When CR and are unequal, B will also change.

A = 0i n <H) + P1 Q 1 « 8.91 x 101 4 (1.0 x 10“5) + 1.432 x 109 « 10.34 x 10S

-1.233 x 1 0 5 +Vl.522 x 1 0 1 0 4(1.034 x l010)(-296)* LMJ = ------------------------------------------------------T7)-----------------------------------

2(1.034 x 10xu) .

[M] = 3.39 x 106/2.068 x 101 0 = 1.638 x 10" 4

Pm « 8.91 x 101 4 x 1.393/1.638 = 7.59 x 101 4 log 0m « 14.880

PlOi = 1.432 x 109(1.393/1.638) = 1.219 x 109 log 01 Q 1 = 9.086

Again refine [M].

A = 31 U <H> + P1 0 1 « 7-59 x 1014(1 x 10'5) + 1.219 x 109 = 8.81 x 109

171

-1.233 x 105 + V 1.522 x 101 0 - 4(8.81 x 109)(-296)2(8.81 x 1 0 9)

[H] = 3.11 x 106/1.762 x 101 0 = 1.765 x 10~ 4

Pm = 7.59 x 101 4 x 1.638/1.765 = 7.04 x 101 4 !°g Pm = 14.847

andP1 Q 1 = 1.219 x 109 x 1.638/1.765 = 1.131 x 109 log ^101= 9’05^

By now one can see that the logarithm of the change in the P

values is a trifle less than half of the previous change. Further

refinement would decrease both P values by about 0.03 log units.

The significant titration points for each metal-polyphosphate

system are listed in table 49. These points occur where the ratio of

equivalents of acid to moles of ligand, defined as "a," is equal to 0 ,

0.5, 0.75, 1.0, 1.5, and 2.0. When the metal solution already contains

acid the "a11 is corrected accordingly, and there are no data for a = 0

on the titration curve. Data for tetramethylamraonium polyphosphates

without metal ions added are also listed for comparison. Unless noted

otherwise, and are the same as listed in figures 31, 32, and 33.

A comparison of the .results obtained here and those of other

workers- is shown in table 50. Perhaps the best comparison is with the

lo® Km m . - 5.89

iog Pl 0 1 = 9.05 - 0.03 = 9.0'2

Discussion

/

172

TABLE 49

TITRATION DATA FOR POLYPHOSPHATES

System•'Recorded pH at Given "a”

0 . 0 0.5 0.75 1 . 0 1.5 2 . 0

r 4n - p 2o 7 10.03 8.84 8.38 7.32 6.18 4.78

Mg-P20 7 9.20 7.08 6.61 6.32 5.68 4.22

Mn-P20?a 9.56 6.65 5.75 4.79 3.87 3.49

Fe(ll)-P207b 8.45 • • • % • • .• • • * * • • • • • • • • •

C o C i i ) - p 2o 7c 8.16 6.36 6.03 5.62 4.62 3.94

Ni-P20?d 7.77 6 . 2 0 5.79 5.53 5.07 3.81

Cu(ll)-P207c,e 6.87 5.52 5.12 4.67 3.95 3.31

Zn-P20?a • • • • 6.33 5.67 4.76 3.92 3.55

A 1-P2°7C • • • • 3.87 3.46 3.16 2.83 2.64

Fe(lll)-P207a » • * • 3.81 3.62 3^42 3.05 2.79

V - P3°10 9.98 8.73 8.23 7.27 5.79 4.50

Mg-P3°ip 9.44 6.82 6.64 6.18 5.27 3.91

Mn"P3°10 8.25 6 . 1 1 5.70 5.24 4.32 3.60

Fe(ll)-P3O1 0 8 . 1 1 6.15 5.69 5.23 4.30 3.37

Co(ll)-P3 0 1 0 8.08 5.99 5.62 5.21 4.38 3.43

Ni"P3°10 7.87 5.99 5.68 5.30 4.52 3.56

Cu(ll)-P3 0 10e 7.15 5.64 5.17 4.60 3.72 3.18

Zii-P3°io • • • • 5.82 5.36 4.92 4.26 3.65

Al-p3°io • • # • 4.00 3.57 3.23 2.80 2.55

Fe(lll)-P301 0 • • • # 4.41 3.89 3.51 2.97 2.65

173

TABLE 49 — Continued

SystemRecorded pH at Given "a"

0 . 0 0.5 ' 0.75 1 . 0 1.5 2 . 0

V ' V l 3 9.01 8 . 1 0 7.74 7.38 6.63 4.64

7.44 6 . 6 8 6.39 6.09 5.41 3.95

Mn- V l 3 C ’f 6.87 6 . 0 0 5.61 5.21 4.48 3.65

Fe(II)-P407f 6 . 6 8 5.84 5.50 5.17 4.48 3.64

Co(lI)-P4On 6.89 5.98 5.61 5.28 4.57 3.60

Hi-P4°i3f 6.91 6.07 5.73 5.39 4.66 3.63

Cu(XI)-P.0,3e 6.19 5.35 4.82 4.46 3.60 3.11

Zn-P4°13 • • • • 5.94 5.56 5.08 4.42 3.83

Al-?4°13f • • • • 4.20 3.67 3.35 2.87 2.61

Pe(III)-P4 013£ • • * • 4.27 3.89 3.61 3.13 2.81

NOTE: Large quantity of precipitate.^Precipitation even when concentrations are cut in half.Slight precipitation.

' All concentrations are doubled.eThe pH of the 0.10F CuCNO^g solution was 2.755. The "a" values are not corrected here for this H . contribution.

fC„ = 0.00360; C. = 0.00341.M : L

174

TABLE 50

COMPARISON OF RESULTS WITH THOSE OF OTHER WORKERS

Complexlog “m l Iofi Sjhl

ThisLaboratory Others This

Laboratory Others

Cu-P2 0 7 9.02 9.07b 5.89 5.37b

Cu-P3°10 8.74 8.70° 5.87 5.69c• * • * 7.85d * * • • 4.34d

Cu-PA°i3 9.49 9.44e 6.67 6.66e«s-r2o 7 5.34 5.41f 3.55 3.06f

M«-P3°10 5.46 5.83f 3.48 3.34f• • • * 5.76d • • • • 3.27d• • • • 5.36s * 9 m m 3.33s

Mg-P4°13 5.88 ' 6.04f ‘ 4.44 3.74fC o - P ^ 6.96 7.36h 4.61 4.07h

Oo-p3°10 7.59 8.13h 4.99 5.16h* • • ♦ 6.98d • • m • 3.81d

Co-pA°i3 7.61 * m • • 5.40 * « * •

Ni-P.,0, 6.47- 7.01h 3.91 3.81hHl-PjOjjj 7.22 7.90h 4.64 5.01h

• • • • 6.78d • • • • 3.65d

N1-P*°13 7.41 • • * • 5.40 . * • • •

NOTE: ^his laboratory, [1 = 1 with tetramethylammonium chloride.

* 175

NOTES FOR TABLE 50 — Continued

COMPARISON OF RESULTS WITH THOSE OF OTHER WORKERS

OTHERS: ^Schupp, Sturrock, and Watters (1963), (1 = 1.0 with.tetramethylammonium nitrate. ' ,

CSturrock, Loughran, and Watters (1962), H = 1.0 with tetramethylammonium nitrate.

^Ellison and Martell (1964), p. « 0.1 with RC1.

Matters and Matsumoto (1966), p = 1.0 with tetramethylammonium nitrate.

£Watters and Machen (1968), p = 1.0 with tetramethylammonium chloride.

®Roppongi and Kato (1962), p c 0.1 with KCi.

^Hammes and Morrell (1964), p s 0.1 with tetramethylammonium chloride.

* * *

copper complexes. Watters and his coworkers used a dropping-amalgam

electrode to determine the concentration of uncomplexed copper(II) ion

directly (Schupp, Sturrock, and Watters, 1963; Sturrock, Loughran, and

Watters 1962; Watters and Matsumoto 1966). Therefore, no iterative

procedure was required to determine CM], The results for copper

triphosphate and tetraphosphate agree closely. In the case of copper

pyrophosphate, a precipitate formed in slightly acid solution when the

pH was less than 4.3. This may have been the neutral species, CuH^P^O^.,

which would be expected to be only slightly soluble and to raise the

acidity of the solution. The constants, and are a^out

log unit too high, just as would be expected if other acid pyrophosphate

complexes form in addition to CuHP^O^”.

The results for nickel and cobalt triphosphates may be compared

with those of two other groups of workers. It is interesting to observe

that the results in this laboratory are about halfway between those of

Hammes and Morrell (1964) and of Ellison and Martell (1964). .The former

authors used an ionic strength of only 0.1M, so that their constants

would be expected to be somewhat higher. Ellison and Martell used 0.1M

KC1 instead, of (CH^)^NCl. Since potassium ions also form weak complexes

with polyphosphates (Watters and Matsumoto 1967), the additional lowering

of the pH by transition metal ions would be less, and thus the calculated

constants would be lower. The constant, f°r cobalt pyrophosphate i

too.high because a precipitate formed in slightly acid solution.

Even when no pyrophosphate precipitate formed, as with ^

magnesium, the constant, appears to be high, possibly because the

method does not take other complexes such as MH^L into account. The

titration curves show that the pH is still slightly depressed a little

beyond the point where a = 2 , and that such complexes may form in small

concentrations.

In the case of magnesium triphosphate.the results of three

other groups of workers (Watters and Machen 1968; Ellison and Martell

1964; and Roppongi and Kato 1962) may be compared. All agree closely

for the acid stability constant, I*1 closest agreement with the

■results obtained in this laboratory for are those of Roppongi and

Kato (1962). Even though magnesium is not a transition metal, experi­

mental work was done for comparison. The ionic radius of the magnesium

ion is just a little less than that of the first row divalent transition

metal ions.

177

Another fact not previously taken into account is that the

third acid constant becomes important below a pH of 4.0. The third acid3

constant enters the y expression as (3C^ - » an<* ^ is aboutu

1% of .the total y expression at pH 4. This additional factor becomes

very important in calculating the constants for the trivalent metals

where the pH is about 3 at a = 1.5.An attempt was made to determine the constant for or 1

?1 2 i directly. The algebraic equations can be developed in the same way

by assuming ^ 1 2 1 important metal-complex constants.

There is a similar quadratic equation for CM], and the slope-intercept

equation is' [0H - ct][M]CH)Pl u + (CH - 2cl)[m](h)2 P121- = (CL - V pon(H) +<2CL " °H)P021(H)Z + <3CL " CM)P031<H) " °H‘

The intercept, b, is (C^ - Cl > [ M ] ( H ) T^e sloPe *-s ^121* T^ese can2be found by plotting (C - 2CT)[M](H) as x, and the entire right side ofH L

the equation as y. Unfortunately, the acidity is so high that the ApH

between the ligand and metal-complex titrations does not give a positive

slope. All the acid phosphates exist in equilibrium. Furthermore,

hydrolysis of the polyphosphates to lower phosphates also does occur. It

follows that the method used here is limited to the titration of complexes

where the pH is in the neutral range.

Other complexes, M 2L and are indicated by the titration

curves when the metal-polyphosphate ratio is other than 1:1. These both

were found in calculations based on data obtained with a dropping amalgam

electrode (Schupp, Sturrock, and Watters 1963; Sturrock, Loughran, and

Watters 1962; and Watters and Matsumoto 1966). In slightly acid

178

s.olution CuHL>2 and CuH^Lg were also found. Above pH 9 the complex,

CuOHL, where the OH group is attached to the metal ion, is also important.

apparently forms as soon as the ferric nitrate solution is added to the

polyphosphate. The precipitates with triphosphate and titraphosphate-

slowly dissolve to form a soluble complex within thirty minutes. Then ■

the titration may proceed, but a soluble hydroxy-polyphosphate complex

is possible. Aluminum forms no visible precipitate, but it probably also

forms hydroxy complexes^, as it is known to do in aqueous nitrate or

chloride solution.

The possibility of hydrolysis of the polyphosphates to lower

phosphates during the titration was also checked. The least stable

polyphosphate used, tetraphosphate, was titrated to a pH of 2.9. Then

the solution was allowed to stand for twenty minutes, and the change in

pH was only 0.004 unit. Hence, the polyphosphates were assumed to bew —

stable during the course of a titration at 25 C. Usually, the titrations

required about thirty minutes, and the solutions were in the acid range

for only a small part of that time.

Hydroxide complexes of iron(lll) and aluminum may be very important since '

these ions begin to hydrolyze at a pH of 2. The iron(lll) hydroxypyro-

phosphate may be pictured asOH 0X O"

H o0 ^ I — — P.

The solubility of ferric hydroxide is so small that some of it

4

. • 179

The remaining stability constants obtained are listed in table

51. As one would expect, the stabilities of the trivalent metal phos­

phates are much greater than those of the divalent metals. The log

iron (III) and aluminum is about two units higher than for copper, which

forms the most stable divalent metal complexes. The stabilities of the

complexes of iron(lll) are only slightly greater than those of aluminum.

This fact suggests that the use of d orbitals for bond formation with

ferric ion is of only slight importance when the ionic charge is high.

It should be emphasized that the stability constants with the trivalent

ions are only approximate, because hydroxy complexes are probably also

formed.

McNabb, Hazel, and Baxter (1968) recently published a list of

stability constants for lanthanum polyphosphates. Their log values

were 4.66, 6.56, and 6.59 for the pyrophosphate, triphosphate and tetra-

phosphate respectively. The log values were 0.85, 2.90, and 3.29.

The ionic strength was adjusted to 0.1 with tetramethylammonium chloride,

but the sodium ion was apparently not removed from the polyphosphate by

/ion exchange. Mass balance equations were listed, but working equations

were not developed.

Malik and Sharma (1968, pp. 29 and 503) characterized some

pyrophosphate precipitates of the trivalent metals, chromium, iron, and

aluminum. The concentrations of iron(lll) and chromium(lll) could be

determined by polarography. With iron as an example, typical complexes-6 +2 -

were Fe2 (p2° 7 ^ 3 » Fe4^P2°7^3* Fe2F2°7 ' ant* FeF2°7 *

180y '

TABLE 51

METAL-POLYPHOSPHATE CONSTANTS

Complex L°s Km l Lo£ ^MHL

M 11-P3 O 1 0 7.59 5.16

Mn"P4°13 7.65 5.59

Fe(lX)-P3 O l 0 7.50 5.15

Fe(ll)-P4 0 13 7.68 5.44

Zn-P3°io 7.70 4.97

Zn“P4°13 8 . 0 0 5.86

A1-P20? 11.65 6.80

A 1 -P3°10 11.13 6.57

A1"P4°13 11.32 7.53

Fe(lIl)-P3 O 1 0 11.32 6.92

Fe<IXl)-P4 0 1 3 11.69 7.81

. ' 181

The trends in stability of the divalent metal complexes are as

follows:For with triphosphate, Mg < Ni < Fe < Co, Mn < Zn < Cu. '

‘ For with triphosphate, Mg < Ni < Co, Zn < Fe, Mn < Cu.

For with tetraphosphate, Mg < Ni < Co, Mn, Fe < Zn < Cu.

' For with tetraphosphate, Mg < Ni, Co Fe < Mn < Zn < Cu.

As is usual with transition metal complexes, those of copper are the most

stable and are followed by those of zinc in most cases. Of the elements,

manganese through nickel, the usual order of stability is reversed. The

complexes of nickel are the least stable, while those of manganese are

the most stable. One must remember that the metal-ligand bond is through

oxygen, and the bond is thought to be chiefly ionic. An explanation for

the stronger copper and zinc bonds is that one of the PO^ groups appears

to be bidentate (Brintzinger and Plane 1967), and thus the whole poly­

phosphate is at least tridentate. The determination of the stability con­

stants of iron(ll) appears to be a new contribution. In this laboratory

the possibility of oxidation of iron(ll) was no problem as long as the

solution was kept under a nitrogen atmosphere before and during the

titration.

The relative stabilities of a given metal ion with the various

polyphosphates are of. some interest. In every case the stability of the

metal tetraphosphate is somewhat greater than the stability of the

triphosphate. This is to be expected simply because there are more oxy­

gen atoms available for bonding to the metal ion. There is also a statis­

tical factor favoring complex formation with tetraphosphate. On the basis

of stability per phosphorus atom, the triphosphates are more stable. The

182

relative position of the pyrophosphate is erratic. Sometimes it is the

most stable; sometimes the least stable. Some of the pyrophosphate

stabilities could not even be determined because of precipitation. This4

was true of manganese, zinc, iron(ll), and iron(lll). The precipitation

was greatest around a pH of 4 where one would expect MHPgO^ or even

neutral MHgPgOy to form. It is well known that neutral complexes are

often insoluble in aqueous solution.Attempts to prove the actual presence of neutral species by

means of solvent extraction or electrophoresis were futile. Sometimes

there was no movement during electrophoresis, but the spot appeared to

be a precipitate, which would not move regardless of the charge.

For solvent extraction the conditions were adjusted so that the

most probable pyrophosphate was either 0 0 2 ^ 2 ^ 7 or a ^6.5 the dicobalt pyrophosphate precipitated and remained as a suspension

in the aqueous layer. At a pH of 4.0, where might form, no

precipitate appeared. Still there was no evidence of extraction by either

methyl isobutyl ketone or 2-octanol. Concentrated hydrochloric acid was

'shaken with the organic solvent to determine if any cobalt had been

extracted. There was no evidence of any blue color in the acid.

Another set of experiments was run in order to see if the two

most important complexes are really ML and MHL. The metal-polyphosphate

ratios were kept at 1 :1 , but the molar concentration of each was varied

from 0.001 to 0.005. If the calculated stability constants by the slope-

intercept method are the same, then other metal complexes can be assumed

to be unimportant. The results for three polyphosphate complexes are

listed in tables 52, 53, and 54. The relative standard deviation is

183

TABLE 52

EFFECT OF CONCENTRATION ON STABILITY CONSTANT CALCULATIONS OF NICKEL PYROPHOSPHATE

CM CL pH at [M] LogKMLa

Logw

a « 0.5 0.75 1.5

0.00099 0.00097 6.52 6 . 2 2 5.51 3.5 x i o " 4 6.25 3.88

0.00193 0.00189 6.42 6.05 5.35 5.6 6.29 3.88

0.00196 0.00192 6.38 6.05 5.31 5.1 6.46 . 4.03

* 0.00365 0.00357 6.26 5.91 5.16 8 . 1 6.49 3.96.

0.00491 0.00480 6 . 2 1 5.80 5.08 11.5 6.47 3.91

NOTE: ^Standard deviation is 0.10.^S.tandard deviation is 0.06.

S

184

TABLE 53

EFFECT OF CONCENTRATION ON STABILITY CONSTANT

CALCULATIONS OF POBALT(II) TRIPHOSPHATE

CM “ CLpH at

£

i—i LogKm l *

LogKMHLba ■= 0.5 0.75 1.5

0.00098 6 . 1 1 5.76 4.69 1 . 2 2 x 1 0 " 4 7.48 5.00

0.00193 6.03 5.68 4.54 1.9 7.56 5.02

0.00385 5.98, 5.94 5.61 4.39, 4.35 2 . 8 7.59 4.99

0.00509 5.91 5.55 4.27 4.! 7.67 5.00

NOTE: ^Standard deviation is 0.07.Standard deviation is 0.01.

TABLE 54

EFFECT OF CONCENTRATION ON STABILITY CONSTANT

CALCULATIONS OF COBALT(ll) TETRAPHOSPHATE

CM CL pH at [M] LogKMLa

Log^ L *

a = 0.5 0.75 1.5

0.00083 0.00083 6.09 5.78 4.89 0.91 x l p -4 7.76 5.73

0.00163 0.00162 6.03 5.73 4.75 1.83 7.71 5.59

0.00327 0.00328 5.96 5.61 4.57 2.5 7.77 5.54

0.00346 0.00328 5.90 5.59 4.54 3.0 7.78 5.59

0.00440 0.00442 5.94 5.60 4.54 3.0 7.84 5.56

NOTE: ^Standard deviation is 0.05.Standard deviation is 0.075.

185

given in logarithm units, and this is equivalent to less than 207.. The

result most likely to be out of line is the first, where the concentra­

tions are 0.001F or less. The acid for the titration was also diluted,

and the pH changes were not as great. There is a very slight trend

toward larger stability constants as the concentrations become greater,;

but this increase is not statistically significant. It may be assumed

for the polyphosphates listed in these tables that under these experi­

mental conditions the only important metal complexes are ML and MHL.

In the presence of an excess of metal ion there is an increased

probability of forming or MgHL as .well. That these do form is

indicated by the fact that different values are obtained for the above

'constants if the metal-polyphosphate ratios are much larger than 1 :1 .

Especially notable is the decrease in the apparent stability constant for

MHL. When the metal-polyphosphate is much less than 1:1, the two terms

in the numerator of the quadratic equation for [M] are practically the.

same, and.the difference is too near to zero to leave any significant

figures.

1 A graph showing the titration of zinc triphosphate at various

metal-phosphate ratios is shown in figure 34. When the ratio is 0.5, the

titration is about what one would expect when ZnL and ZnHL are half

formed, and the other half is triphosphate ion. Going beyond the ratio

of 1.0 continues to depress the pH, but apparently the formation of M 2L

is not quantitative. No calculations for the stability of M^L were

undertaken. Addition of still more metal ion beyond a ratio of 2:1

produced a precipitate.

. FIGURE 34

TITRATION OF VARIOUS RATIOS OF Zn AND TRIPHOSPHATE

0 . 2 0 millimole of triphosphate

(Zn:P^O^Q ratio as indicated)

-\

‘••.2.0 X

2 4 6 ml HN03 ' 0 100N

187

Summary

A mathematical'method of determining the stability constants

of metal polyphosphates from a potentiometric titration has been utiliz­

ed. The complexes of magnesium, aluminum, and several transition metals

with pyrophosphate, triphosphate, and tetraphosphate were studied. The

stability constants of the divalent transition metals with the poly-7 9phosphates ranged between 10 and 10 . In the case of aluminum and

iron(lll) the stability constants are greater than 10^. When there is

one hydrogen attached to the polyphosphate group, the stability constants5 7are about 1 0 for the divalent metals and 1 0 for the trivalent ones.

REFERENCES

Addison, C. C. , and Sutton, D.1964. J. Chem. Soc. 1964:5553.

Angell, C. A., and Gruen, D. M.1967. J. Inorg. Nucl. Chem. 29:2243.

Awad, W. I., Gawargious, Y. A., Hassan, S. S. M., and Milad, N. E1966. Anal. Chim. Acta. 36:339.

Ball, R. G., Manning, D. C., and Menis, 0.1960. Anal. Chem. 32:621.

Ballaux, C., Dams, R., and Hoste, J.1967. Anal. Chim. Acta. 37:164.

Ballhausen, C. J., and Jorgensen, C. K. 1955. Acta Chem. Scand. 9:397.

Barney, J. E., and Bertolocini, R. J.1957. Anal. Chem. 29:1187.

Beamish, F. E.1966. The Analytical Chemistry of the Noble Metals. New York

Pergamon Press, p. 57.

Belcher, R., and Fildes, J. E.1961. Anal. Chim. Acta. 25:34.

Belcher, R., and Ingram, G.1952. Anal. Chem. Acta. 7:319.

188

189

Belcher, R., MacDonald, A. M. G., and Nutten, A. J.1954. Mikrochim. Acta. 1954:104.

Bergman, J. G., and Cotton, F. A.I960. Inorg. Chem. _5:l420.1966. Inorg. Chem. 5_:1208.

Bergman,'J. G., and Martin, R. L.1962. Anal. Chem. 34:911.

Bergman, J. G., and Sanlk, J.1957. Anal. Chem. 29:241.

Bertrand, J. A., and Plymale, D. L.1964. Inorg. Chem. _3:775.1966.- Inorg. Chem. _5:879.

Biagetti, R. V., and Haendler, H. M.1966. Inorg. Chem. 5:383.

Bjerrum, J.1944. Kg. Danski Videnskob Selskab Mat. Fys. Medd., No. 4. 21.

Blake, A. B., and Cotton, F. A.1963. Inorg. Chem. Ji:906.

Boltz, D. F. •1958. Colorimeteric Determination of Nonmetals. New York:

Interscience, chap. 5.

Brintzinger, J., and Plane, R. A.1967. Inorg. Chem. £:623.

Buffagni, S., and Dunn, T. M.1960. Nature. 188:937.

Carlson, G. A., McReynolds, J. P., and Verhoek, F. H.1945. J. Am. Chem. Soc. 67:1334.

190

Chang, J. C., and Garner, C. S. 1965. ■ Inorg. Chem. _4:209.

Chapman, F. W. , and SherwoodtJ R. M.1957. Anal. Chem. 28:172.

Cheng., F. W.1959. Mlcrochemlcal Journal. 3:537.

Clark, R. J. H., and Williams, C. S.1965. Inorg. Chem. 4:350.

Coleman, J. S.1966. J. Inorg. Nucl. Chem. 28:2371.

Collman, J. F.1968. Stanford University. Lecture at The Ohio State'University,

May 1968, at Columbus, Ohio.

Cosgrove, J. F., Bastian, R. P., and Morrison, G. H.1958. Anal. Chem. 30:1872.

Cotton, F. A., and Dunne, T. G.1962. J. Am. Chem. Soc. 84:2013.

Cotton, F. A., Dunne, T. G., and Wood, J. S.1965. Inorg. Chem. 4:318.

Cotton, F. A., and Elder, R. C.1965. Inorg. Chem. ^:1145.1966. Inorg. Chem. 5:423.

Cotton, F. A., Faut, 0. D., and Mague, J. T.1964. Inorg. Chem. 3:17.

Cotton, F. A., and Gooagame, M.1961. J. Am. Chem. Soc. 83:1777.

Cotton, F. A., Goodgame, D. M. L . , and Goodgame, M.1961. J. Am. Chem. Soc. 83:4161 and 4690.

Cotton, F. A., Goodgame, D. >1. L., Goodgame, M. , and Haas, T. E.1962. Inorg. Chem. 1:565. (Cotton et al. 1962)

Cotton,- F. A., Goodgame, D. M. L., Goodgame, M., and Sacco, A.1961. J. Am. Chem. Soc. 83:4157. (Cotton et al. 1961)

Cotton, F. A., Goodgame, D. M. L., and Soderburg, R. H.1963. Inorg. Chem. _2:1162.

Cotton, F. A., and Weaver, D. L.1965. J. Am. Chem. Soc. 87:4189.

.Crocket, D. S., and Grossman, R. A.1964. Inorg. Chem. 3:644.

Douglas, B. E., and McDaniel, D. H.1965. Concepts and Models of Inorganic Chemistry. New York:

Blalsdell, p. 121.

Dubsky, J. V. , and Trtilek, J. 1933. Mlkrochemle. 12:315.

Einstein, F. W. B., and Rodley, G. A.1967. J. Inorg. Nucl. Chem. 29:347.

El-Awady, A. A., Bounsall, E. J., and Garner, C. S.1967. Inorg. Chem. 6:79.

Ellison, H . , and Martell, A. E.1964. J. Inorg. Nucl. Chem. 26:1555.

Elshelmer, H. N., Johnston, A. L. , and Kochen, R. L.1966. Anal. Chem. 38:1684.

Eswien, R. P., Howald, E. S., Howald, R. A., and Kuton, D. P. 1967A. J. Chem. Soc. 1967A:437.

192

Everett, G. W. , and Holme, R. H.1965. J. Am. Chem. Soc. 87:5226.

Filby, R. H.1964. Anal. Chim. Acta. 1,31:437.

Fildes, J. E., and MacDonald, A. M. G.1961. Anal. Chim. Acta. 24:121.

Fine, D. A.1962. J. Am. Chem. Soc.; 84:1139.1965. Inorg. Chem. It: 345.

Garaka, K. J.1968. Anal. Chem. 40:809.

Gerlach, J. L., and Frazier, R. G.1958. Anal. Chem. 30:1142.

Gill, J. B., and Riaz, S. A.1969A. J. Chem. Soc. 1969A:183.

Gill, N. S., and Nyholm, R. S.1959. J. Chem. Soc. 1959:3997.

Good, M. L., and Srivaatava, S. C.1965. J. Inorg. Nucl. Chem. 27:2429.

Goodgame, D. M. L., and Goodgame, M.1963. J. Chem. Soc. 1963:207.

Griffith, E. J.1964. Inorg. Chem. 26:1381.

Griffith, E. J., and Buxton, R. L.1965. Inorg. Chem. j4:549.1967. J. Am. Chem. Soc. 89:2884.

193

Harames, 6 . C., and Morrell, M. L.1964. •j. Am. Chem. Soc. 86:1497.

Berber, R. H., and Irvine, J. W*1956. J. Am. Chem. Soc. *^8:905.1958. J. Am. Chem. Soc.. 80:5622.

Holm, R. H., and Cotton, F. A.1959. J. Chem. Phys. 31:788.

Hull, 0. E., and Gilmore, J. T.1964. Anal. Chem. 36:2072.

Iwasaki, I., Utsumi, S., and Ozawa, T.1952. Bull. Chem. Soc., Japan. 25:226.

Job, P.1928. Ann. Chim. [10]. 9^:113.

Jorgensen, C. K.1954. Acta Chem. Scand. J3:1495.

Katzin, L. I.1952. J. Chem. Phys. 20:1165.1954. J. Am. Chem. Soc. 76:3089.1961. J. Chem. Phys. 35:467.1962. J. Chem. Phys. 36:3034.

Katzin, L. I., and Gebert, E.1950. J. Am. Chem. Soc. 72:5451 and 5455.1953. J. Am. Chem. Soc. 75:2830.

King, H. C. A., Kor'os, E., and Nelson, S. M.1963. J. Chem. Soc. 1963:5449.1964. J. Chem. Soc. 1964:4832.

Kolthoff, I. M . , and Willraan, A.1934. J. Am. Chem. Soc. 56:1007.

194

Kraus, K. A., Nelson, F., Clough, F. B., and Carlson, R. C.1955. J. Am. Chem. Soc. 77:1391.

Kristjanson, A. M . ,• and Lederer, M. .1959. Journal of Less-Common Metals. _1:245.

Laltlnen, H. A., and Lin, Zul-flng.1963. Anal. Chem. 35:1405.

Lawson, D., and Wilkinson, G.1965. J. Chem. Soc. 1965:1900.

.Lederer, M.1958. J. of Chromatography. 1:279.

Llndenbaum, S., and Boyd, G. E.1963. J. Phys. Chem. 67:1238.

Lister, M. W . , and Rosenblum, P.1960. Can. J. Chem. 38:1827.

MacNevin, W. M., and Crummett, W. B.1954. Anal. Chim. Acta. 10:323.

MacNevin, W. M., and Dunton, M. L.1957. Anal. Chem. 29:1806.

MacNevin, W. M., and McKay, E. S.,1957. Anal. Chem.■ 29:1220.

Machen, R. C. -1967. Ph.D. dissertation, The Ohio State University at Columbus.

Magor, B. S., and Smith, T. D.1968A. J. Chem. Soc. 1968A:1753.

Malik, W. U . , and Sharma, C.. L.1968. J. Indian Chem. Soc. 45:29 and 503

Malmstadt, H. V., Fett, E. R., and Winefordner, J. D.1956. Anal. Chem. 28:1878.

Matwiyoff, N. A.,1966. Inorg. Chem. 5:789.

McNabb, Hazel, J. F., and Baxter, R. A.1968. J. Inorg. Nucl. Chem. 30:1585.

Miller, D. L., and Westheimer, F. H.1966. J. Am. Chem. Soc. 88:1507.

Moore, G. E., and Kraus, K. A.1952. J. Am. Chem. Soc. 74:843.

Moore, T. E., Gootman, E. A., and Yates, P. C.1955. J. Am. Chem. Soc. 77:298.

Moore, W. J.1962. Physical Chemistry. 3rd ed. Englewood Cliffs, N. J.:

Prentice-Hall, p. 351.

Morris, D. F. C., Reed, G. L., Short, E. L., Slater, D. N., and Waters, D. N.

1965. J. Inorg. Nucl. Chem. 27:377.

Orgel, L. E.1955. J. Chem. Phys. 23:1004.

Osterheld, R. K., and Langguth, R. P.1955. J. Phys. Chem. 59:76.

0ye, H. A., and Gruen, D. M.1964. Inorg. Chem. <3:836.1965. Inorg. Chem. 7:1173.

Pasternack, R. F., and Plane, R. A.1965. Inorg. Chem. 4:1171-.

Pauling, Peter*1966. Inorg. Chem. 5,: 1498.

Paulsen, I. A., and Garner, C. S.1962. J. Am. Chem. Soc. 84:2032.

Pflaum, R. T., Frohliger, J. 0., and Berge1962. Anal. Chem. 34:1812.

Qulmby, 0. T.1954. J. Phys. Chem. 58:614.

Ramette, R. W.1963. J. Chem. Ed. 40:71.

Raspanti, G.1967. Z. Anal. Chem. 225:24.

Rodley, G. A., and Smith, P. W. 1967A. J. Chem. Soc. 1967A:1580.

Roppongi, A., and Kato, T.1962. Bull. Chem. Soc. Japan. 35:1086

Rowe, R.1965. Anal. Chem. 37:368.

Rudolph, J. S., and Nadalin, R. J.1964. Anal. Chem. 36:1815.

Rutner, E.1961. J. Phys. Chem. 65:1027.

Sato, T.1967A. J. Chem. Soc. 1967A:547.

Scaife, D. E., and Wood, K. P.1967. Inorg. Chinn,. b:358. •

Schbniger, W.1955. Mikrochim. Acta. 1955:123.1956. Mikrochim. Acta. 1956:869.

Schulz, I.1956. Z. Anorg. Allgem. Chem. 287:106.

Schupp, 0. £., Sturrock, P. E., and Watters, J. X.1963. Inorg. Chem. 2:106.

Shen, C. Y., Stahlheber, N. E., Dyroff, D. R.1969. J. Amer. Chem. Soc. 91:62.

Shukla, S. K.1958. J. of Chromatography. _1:457.1959. Journal of Less-Common Metals. 1:333.

Shukla, S. K., and Lederer, M.1959. Journal of Less-Common Metals. 1:255.

Smithson, J. M. , and Williams, R. J. P.1958. J. Chem. Soc. 1958:457.

Stern, M., Shwachman, H., Licht, T. S., and Bethune1958. Anal. Chem. 30:1506.

Sturrock, P. E., Loughran, E. D., and Watters, J. I1962. Inorg. Chem. 1:457. '

Thilo, E., and Ratz, R.1949. Z. Anorg. Chem. 260:255.

Vallarino. L. M.1965. Inorg. Chem. ^:161.

VanLoon, J. C.1968. Analyst. 93:788.

Viebock, F.1932. Ber. 65:493.

Vosburgh, W. C. , and Cooper, G. G. 1941. J. Am. Chem. Soc. 63:437.

Watters, J. X., and Machen, R.1968. J. Inorg. Chem. 30:2163.

Watters, J. I., and Matsumoto, S.1964. J. Am. Chem. Soc. 86:3961.1966. •• Inorg. Chem. 5^:361.1967. J. Inorg. Nucl. Chem. 29:2955.

Watters, J. I., and Simonaltis, R. A.1964. Talanta. 11:247.

Watters, J. I., Sturrock, P. E., and Simonaltis, R. A.1963. Inorg. Chem. 2:765.

Weed, E. D.1964. Ph.D. dissertation, The Ohio State University at Columbus.

West, P. W., and Coll, H.1956. Anal. Chem. 28:1834.

Westman, A. E. R., and Scott, A. E.1951. Nature. 168:740.

Whiteker, R. A., and Davidson, N.1953. J. Am. Chem. Soc. 75:3081.

Willett, R. D., and Liles, D. L., Jr.1967. Inorg. Chem. 6:1666.

Wolsey, W. C., Reynolds, C. A., and Kleinberg, J.1963. Inorg. Chem. 2:463.

Wormser, Y.1948. Bui. Soc. Chim. France. 1948:395.

Wu, C., Witonsky, R. J., George, P., and Rutman, R. J.1967. J. Am. Chem. Soc. 89:1987.