You are on page 1of 10

Stability and solubility relationships between some solids

in the system Pb0-C02-HzO1


PETERTAYLOR AND VINCENT J. LOPATA
Research Chemistry Branch, Atomic Energy of Canada Limited, Whiteshell Nuclear Research Establishn~ent,
Pinawa, Man., Canada ROE ILO
Received May 9, 1983

PETERTAYLORand VINCENT J. LOPATA.Can. J . Chem. 62, 395 (1984).


We have determined equilibrium conditions for each reaction in the solid interconversion sequence PbO =
PbloO(OH)6(C03)6(plumbonacrite) E Pb3(OH)Z(C03)Z =
(hydrocerussite) PbC03 (cerussite), in aqueous carbonate solutions
near 25°C. Our data yield the following Gibbs energies of formation from the elements, bfG0(298.15 K): plumbonacrite,
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

* *
-53 10 3 1 kJ mol-I; hydrocerussite, - 1705 1 1 kJ mol-I; cerussite, -628.0 ? 3.9 kJ mol-' . Corresponding estimated
*
values of ArG0(373 K) are -5120 2 50, - 1643 18, and -612 * 9 kJ mol-', respectively. We have demonstrated the
existence of a narrow stability field for plumbonacrite, and convenient small-scale preparations of hydrocerussite and plum-
bonacrite are described. Gibbs energy data for the solids have been combined with selected data for aqueous lead species in
a detailed computation of solubility behaviour for this system at 25°C.

PETERTAYLORet VINCENT J. LOPATA.Can. J. Chem. 62, 395 (1984).


Nous avons determine les conditions d'tquilibre de chaque reaction dans les interconversions suivantes a l'ttat solide: PbO
= PbloO(OH)6(C03)6(plumbonacrite) S Pb3(0H)Z(C03)2 =
(hydrocerussite) PbCO3 (cerussite), dans des solutions aqueuses
de carbonate i environ 25°C. Nos donntes fournissent les energies de Gibbs suivantes pour la formation i partir des Cltments
*
AfG0(298,15 K): plumbonacrite, -53 10 31 kJ mol-'; hydroctrussite - 1705 ? 1 1 kJ mol-I; cCrussite -628,O 2 3,9 kJ
mol-'. Les AfG0(373 K) sont respectivement tvaluts comme ttant suivantes; -5120 * 50, - 1643 & 18 et -612 ? 9 kJ
mol-'. Nous avons dtmontre l'existence d'un champ ttroit de stabilitt pour le plumbonacrite et nous dtcrivons une preparation
convenable a faible tchelle de l'hydrocerussite et du plumbonacrite. On a combine les donnees d'tnergie de Gibbs des solides
Can. J. Chem. 1984.62:395-402.

avec des donntes choisies pour des composts du plomb en solution aqueuse dans un calcul detail16 du comportement de ce
systeme en ce qui a trait a la solubilitC 5 25°C.
[Traduit par le journal]

Introduction ronmental and mineralogical behaviour of lead (14- 18). We


Many solids exist in the system PbO-C02-H20. Of these, have studied interconversions among the five named solids to
two polymorphs of PbO (litharge and massicot) and three car- determine equilibrium constants and hence t o compute solu-
bonates (cerussite, PbC03; hydrocerussite, Pb3(OH)2(C03)2; bility and stability relationships.
plumbonacrite, PbloO(OH)6(C03)6)occur in aqueous systems
near ambient temperature and pressure (1, 2). Garrels and Experimental
Christ (3) have discussed the stability fields of PbO, cerussite, Starting materials were reagent-grade PbO. KOH, KzCO3 (Fisher),
and hydrocerussite in the system Pb-0,-C0,-H,O at 25°C. and CO, (Matheson). Carbonate-free KOH solutions (J. T. Baker)
Recently, plumbonacrite was reported (4) to be metastable with were used in key determinations. Potassium salts were used in prefer-
respect to decomposition according to reaction [I]: ence to sodium salts to avoid precipitation of NaOH -2PbC03, which
is less soluble than KOH.2PbC03 (19). The PbO reagent was a
[I] PbloO(OH)6(CO3)6 + PbO + 3Pb3(OH)z(CO3)2 mixture of litharge and massicot; this was adequate for most purposes,
since their Gibbs energies of formation differ by only about 1 kJ mol-'
-
Several phases of the type a P b C 0 3 bPbO occur during thermal (20, 21). Samples of litharge and massicot were collected whenever
decomposition of the named carbonates, but d o not normally they were obtained as pure reaction products. Recrystallized litharge,
occur in aqueous systems (5-7). Grisafe and White (7) deter- prepared in an earlier study (22), was also used in a few experiments.
mined the stability fields of these phases in the system
PbO-CO, at elevated temperatures and pressures. A hydrated ( a ) Preparation of cerussite
lead(I1) oxide, 3PbO-H,O or Pb60,(0H),, is also known Pure cerussite was readily made by bubbling COZ through a sus-
pension of PbO (20 g) in dilute acetic acid mol dm-3, 200 cm3)
(8- 1 l ) , and early solubility measurements indicate that it may
at room temperature for 6 h. Unlike Olby (2), we did not obtain
b e marginally stable with respect to reaction [2] at ambient plumbonacrite or hydrocerussite as pure intermediates, although they
temperature (1 I): did occur sequentially as major phases. Our preparation was on a
smaller scale than Olby's, with a more dilute suspension of PbO.

Our interest arose from the possible use of lead compounds ( b ) Preparation of hydrocerussite
to precipitate I4CO3 species, which are radioactive con- Pure hydrocerussite was prepared by shaking a suspension of ce-
russite (5 g) in a solution of KzCO3 (8 g in 200 cm3) for 48 h at room
taminants produced in the heavy-water moderator systems of
temperature. Alternatively, cerussite (5 g) was added to aqueous KOH
Canadian nuclear reactors (12). The formation of lead carbon- mol dm-3, 200 cm3), and the pH of the stirred suspension was
ates may also be important in high-level radioactive waste maintained at 11 for 24 h, using an automatic titrator charged with
disposal, since lead has been proposed as an investment mate- KOH solution (0.1 mol dm-3).
rial for used nuclear fuel in a deep underground disposal vault
(13). Lead carbonates are also of some importance in the envi- ( c ) Preparation of plumbonacrite
Hydrocerussite (5 g) was suspended in a solution (200 cm3) of
KZC03(13 g) and KOH (2 g), which was then stirred for 48 h at
' Issued as AECL-7974. 100°C. The filtered product comprised plumbonacrite (>80%) and
CAN. J . CHEM. VOL. 62, 1984

TABLE1. Solution data for final series of hydrocerussite/cerussite equilibration runs (a) 5 days,
(b) 13 days

Final [C,] x 10' Final HCO,


Initial (mol dm-') pH(meter) pH(calcd)$ (%)t§
HCO, p(cOz)ll
Run* ( 7 (a) (b) (0) (b) (a) (b) (0) (b) (Pa)
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

*Initial [C,] = 3 x lo-' mol dmd3.


iPercentage of total carbonate concentration.
$Calculated from titration data.
I Calculated from average of meter and calculated pH values.
IICalculated from average of all pH measurements and calculations.

hydrocerussite. This was crushed using a pestle and mortar, and re- (3, 23). T h e anionic activity quotient, {OH-)'/{co:-), ob-
turned to the soIution for a further 24 h at 100°C. Care was needed to tained from equilibria such as [6], is sometimes more
maintain the pH near 13, to remain within the narrow stability field of convenient.
plumbonacrite (see below). This procedure was repeated until hydro-
cerussite was reduced to %3 - (usually two or three cycles); it was [6] Pb,(OH)z(C0,)2(s) + cO:-(aq) *
3PbC03(s) + 20H-(aq)
difficult to eliminate completely. The p ( C 0 2 ) value associated with a solution is conveniently
( d ) Interconversion reactions manipulated between and l o 2 Pa by adjusting pH and [C,].
Can. J. Chem. 1984.62:395-402.

Over 200 experiments were run, in which the various solids were Total carbonate activity {C,}, proton activity {H'), and p ( C 0 2 )
exposed to solutions of different pH and total carbonate concentration, are interrelated at 25°C by expression [7], which is derived
[C,].2Most experiments were run in screw-topped plastic bottles; long from equations given by Stumm and Morgan (ref. 23, chapt. 4 )
equilibration runs were performed in flame-sealed Pyrex glass vials. and accepted Gibbs energy data for water and carbonate species
Solid products were identified and their purity estimated by X-ray (e.g. 21, 24):
powder diffractometry (XRD), using CuK, radiation and a Philips
PW-1150 diffractometer equipped with a diffracted-beam mono-
chromator. When necessary, solutions were analyzed for lead by
atomic absorption spectrophotometry, and for carbonate by titration or
using a C02-sensingelectrode (Orion model 95-02). Measurements of x 3.4 x mol kg-' Pa-' (= mol dm-3 Pa-')
pH were made with a glass electrode. When precise pH measurements
were needed (see Table l), the electrode was calibrated with a fresh W e have not distinguished between molal a n d molar concen-
sodium borate buffer solution having pH = 9.00 or 9.18. tration units, since the resulting errors are much less than the
experimental uncertainties. W e have used expression [8] as
Results and discussion proposed by Davies (ref. 2 3 , pp. 134- 136; ref. 25) to estimate
( a ) Theoretical considerations the mean ionic activity coefficient, f,, and hence bicarbonate
The solid interconversion reactions are conveniently ex- and carbonate activities, from the ionic charge, z, and the ionic
pressed by reactions [3] to [5] below. In reaction [3], dis- strength, I .
tinction should be drawn between reactions involving lGharge
and massicot. [8] -log f, = 0.5z2 0.311

[3]
5
iPbO(s)
1
+ C02(g) + ?H20(I) % ; P ~ ~ ~ o ( o H ) ~ ( c o ~ ) ~ ( s )Where necessary, I was calculated from measured pH and [C,]
values by an iterative procedure.
( b ) Reactions of cerussite with KOH
Reactions between cerussite (0.25 g) and K O H solutions (40
cm3; various concentrations up to 5 mol dm-3) were in-
vestigated at room temperature (22 * 2°C). Cerussite alone
was recovered from reactions with final pH below about 9.5.
Hydrocemssite was the major product in the p H range 9.5-12.
Although these reactions involve many aqueous species, the
Plumbonacrite and hydrocemssite occurred together at pH =
equilibrium constant for each reaction can be expressed simply
12-13.5. Lead(I1) oxide w a s formed at pH = 13.5-14; both
as the reciprocal of a carbon dioxide partial p r e ~ s u r ep(COz)
,~
litharge and massicot occurred, often one or the other as a pure
phase, but with no obvious pattern of occurrence. Above pH =
'In this account, [C,] and {C,) represent the sums of the concen-
trations and activities, respectively, of non-complexed carbonate spe- 14, clear undersaturated lead solutions were formed. These
cies, i.e. "H2C03", HCO,, and c ~ : - , and do not include complexes observations are consistent with McIntyre's stability calcu-
such as P ~ C O and~ P~(co~):-. lations reported by Garrels and Christ (3), except for the occur-
Strictly speaking, this term should be fugacity, rather than partial rence of plumbonacrite. Further experiments were designed t o
pressure, but the two are indistinguishable at the pressures of concern define the stability fields more exactly, with special attention t o
here. plumbonacrite.
TAYLOR AND LOPATA 397

CEWSSITE --) I 1 perature of 22 * 2"C, since the equilibrium quotient,


[co:-]/[c,], changes only slightly between 22 and 100°C (see
below).
Our determination of p, is in fair agreement with the value

4
of 27 Pa obtained from the Gibbs energy data recommended by
Sadiq and Lindsay (21). The discrepancy corresponds to a
disagreement of about 1.8 kJ mol-I in the Gibbs energy of
reaction for reaction [5]. Our result is also consistent with
Garrels' estimate of p3 = 10 Pa, which was based on de-
4
0 0.4
ductions from the natural occurrence of cerussite and hydro-
-
.-, cerussite ( 17).
0.2 -
(d) Interconversions involving plurnbonncrite
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

0
Preliminary experiments indicated a narrow stability field for
HYDROCERUSSITE --, plumbonacrite, so subsequent work concentrated on con-
0.5 0.6 0.7 0.8 0.9 1.0 firming this observation and defining the boundaries of the
field.
Plumbonacrite persisted unchanged for seven days in solu-
FIG. 1 . Summary of X-ray diffraction data for products of reactions tions with [C,] = 0.125 rnol dm-3 and [KOH] between 0.02 and
between cerussite (0) or hydrocerussite(o) w~thsolutions having [C,] 0.4 mol dm-3. It was converted to massicot at [KOH] 2 0.6
= [HCO,] + [co:-] = 0.125 mol dm-', for 48 h at 22OC. I(3.50) rnol dm-3, with partial conversion occurring at 0.5 rnol dm-'.
and I(3.29) are the peak heights of the most intense features in the In similar experiments starting with PbO, plumbonacrite was
cerussite and hydrocerussite diffraction patterns (3.50 A and 3.29 A, formed at [KOH] = 0.1 , but not at 0.5 mol dm-'. Slow, partial
respectively).The ordinate function is thus an approximate measure of conversion of plumbonacrite to hydrocerussite occurred at
the extent of conversion. [KOH] 5 0.01 rnol dm-3, and the reverse reaction was ob-
served in some runs with KOH concentrations between 0.05
Can. J. Chem. 1984.62:395-402.

(c) Cerussite-hydrocerussite interconversion and 0.08 rnol dm-'. The latter reaction, accelerated by heating
Experiments in which either cerussite or hydrocerussite was to 100°C, provided the basis for our synthesis of plum-
exposed to carbonate/bicarbonate solutions clearly showed bonacrite, described above. In triplicated experiments starting
that this interconversion is readily reversible at room tem- with mixtures of plumbonacrite and hydrocerussite at [KOH] =
perature. XRD data, summarized in Fig. 1, indicated that hy- 0.01 and 0.1 mol dm-3, the relative amount of plumbonacrite
drocerussite and cerussite were in equilibrium at [coS-]/[c,]= in the product was consistently higher at the higher KOH con-
0.70 + 0.02 and [C,] = 0.125 mol dm-3. At lower [coS-]/[c,] centration. Although these interconversions are less clear-cut
ratios, hydrocerussite was converted to cerussite, and the re- than the hydrocerussite-cerussite reaction, w e could never-
verse reaction occurred at higher ratios. Although equilibrium
conditions are thus narrowly defined, ionic activities could not
theless observe reversible reactions in the sequence PbO =
plumbonacrite G hydrocerussite, thus confirming that plum-
be estimated sufficiently well to calculate p3 as precisely as we bonacrite has a narrow stability field. This narrowness can
wished. Further experiments were therefore run at lower [C,], account for the scarcity of plumbonacrite in nature (4).
down to lo-' rnol dm-3. At these lower concentrations, the Further experiments in sealed glass vials were run for 7- 14
quantity of hydrocerussite or cerussite consumed or formed in days with mixtures of PbO and plumbonacrite (0.3 g each) in
the reactions was too small to determine equilibrium conditions 20 cm3 aliquots of solutions with various KOH concentrations
by XRD. We therefore deduced the direction of reaction, and at [K2C03]= lo-', lo-', and rnol dm-'. Most runs did not
judged whether equilibrium had been achieved, by careful reach equilibrium, but the direction of reaction [9] (equivalent
measurement of pH and [C,]. to reaction [3]) could be determined from solution analyses for
Twenty-five experiments with [C,] between and lo-' carbonate and/or hydroxide.
mol dm-' yielded values for p3 between 8 and 16 Pa. In the
final determination, duplicated experiments were run in sealed [9] $ PbO + C O ~ -+ 23 Hz0 = 61 PbIuO(OH),(CO,)i+ 20H-
Pyrex glass vials, starting with a mixture of cerussite and hy-
drocerussite (0.5 g each) in solutions with various [HCO,]/[C,] 'The results are presented in Fig. 2. From these we estimate the
ratios and [C,] = 3 X lo-' rnol dm-'. This concentration equilibrium constant {OH-}'/{co:-} = lo0.' '0.6 mol dm-3,
provides the best compromise between the respective lim- which corresponds to p , = 10-4.9'0.6 Pa. We did not discern
itations of the carbonate analyses and the estimation of activity reproducible differences between reactions involving PbO re-
coefficients. One series of experiments ((a) in Table 1) was run agent and our synthetic litharge, except that the latter material
for 5 days, and the second ((6) in Table 1) for 13 days. The pH reacted more lethargically. Based on the free energies of for-
was measured immediately after opening each vial. The solu- mation of rnassicot and litharge, p l values for the two forms of
tions were then filtered and analyzed for carbonate and bicar- PbO should differ by a factor of about 1.5, whereas we estimate
bonate by titration; a pH value was then calculated from the an error of a factor of four in p , .
titration data. Most measured and calculated pH values agreed Interconversion between hydrocerussite and plumbonacrite
within 0.04 units, the largest discrepancy being 0.08 units. We was too slow at [C,] 5 lo-' rnol dm-' to pin-point the equi-
conclude that solution analyses were not substantially affected librium at low ionic strength. The close crystallographic simi-
by exchange of C 0 2 with the atmosphere. The results are sum- larity between these phases may promote epitaxial crys-
marized in Table 1. The values of p, at 25OC, determined from tallization with consequent passivation of one phase by the
these runs, range from 12.8 to 14.1 Pa, with an average of 13.3 other. Furthermore, their solubility behaviour is very similar
Pa. No correction was made for the actual experimental tem- (see Fig. 5), so the driving force for interconversion is slight.
398 CAN. J. CHEM. VOL. 62, 1984

TABLE2. Gibbs energy data used to determine equilibrium


relationships

A G:, 298
Species* (kJ mol-') Reference
PbC03(s) -628.0 This work
Pb3(OH)2(C03)2(s) - 1705 This work
. , P ~ I o ~ ( O H ) ~ ( C O ~ )-53
~ ( S10
) This work
I
o Plumbonocrife -PbO observed PbO(s) (litharge) -189.3 21
PbO --Plumbonocrile observed co2(g) -394.38 21
o No observable conversion
H20(1) -237.18 21
"H,C03(aq)" -623.21 21
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

HCO,(aq) -586.89 21
fi (mol dm-3)"2 co:-(aq) -527.94 2I
FIG. 2. Summary of conditions under which interconversion of Pb" (aq) -24.69 21
PbO and plumbonacrite was observed. The numbering on the figure PbOH'(aq) -217.94 21
indicates carbonate concentration (mol dm-3). The bold lines corre-
spond to log K = 0.1 + 0.6 for eq. [9], using eq. [8] to estimate ~b(~~):(aq) -397.73 21
deviation from ideality. Flagged data points correspond to deter- Pb(OH),(aq) - 575.89 21
minations by visual observation ([co:-] = 0.01 mol dm-3) or X-ray Pb20H3+(aq) -250.04 21
diffraction ([co:-] = 0.125 mol dm-'). The remaining observations
are based on solution analyses. pb,(~~):+(aq) -886.42 21
Pb4(0~):+(aq) -928.22 21
Based on the results obtained at [C,] = 0.125 mol dm-, we pb6~(o~)l+(aq) - 1559.59 2 1 , 38, 29, see text
estimate that p, exceedsp, by one to three orders of magnitude, P ~ C O ~ ) -593.37 44, 45
Can. J. Chem. 1984.62:395-402.

so we assign the value p, = Pa. Since plumbonacrite 44, 45


~b(~03)i-(aq) -1141.17
and hydrocerussite differ only slightly in composition, calcu-
lated Gibbs energies and solubilities are much less sensitive to *Our grounds for omission of the following species are presented in
errors in p, than in p , or p,, so precise measurement of p, is the text: PbO(s) (massicot), Pb(OH),(s), Pb60,(OH)4(s), P~,(oH):+-
relatively unimportant in the overall description of the system, (aq), Pb,(OH):(aq), pb6(OH)it(aq).
except in determining the plumbonacrite-hydrocerussite sta-
bility boundary (see expressions [I21 and [13] below). Our experimental Gibbs energy for cerussite, -628.0 r 3.9
kJ mol-I, lies within the range of reported values cited by Sadiq
(e) Gibbs energies of formation of the carbonates and Lindsay (21), -625.5 to - 629.0 kJ mol-I, and is not
When expression [lo] is applied successively to reactions [3] significantly different from their recommended, recalculated
to [5], using Gibbs energy data from ref. 21, we obtain re- value of -629.7 kJ mol-I, originating from the calorimetric
lationships [I I] to [13]. These permit us to calculate the Gibbs work of Adami and Conway (26). Our value of - 1705 -+ 11
energies of formation, ArG0(298.15 K), of the three lead car- kJ mol-I for hydrocerussite agrees with their calculation of
bonates from the p(C0,) data.4 - 17 11.6 kJ mol-I , derived from the pioneering solubility
study by Randall and Spencer, the only source they cite (1 1,
21). Our estimate of -5310 ? 31 kJ mol-I for plumbonacrite
+ RT In p(C02, equilibrium, atmospheres) is much more negative than the value of -5121 kJ mol-'
reported by Haacke and Williams (4). The source of so large a
[I 11 ArGO(plumbonacrite, s, 298.15 K) discrepancy is not readily apparent, since few details are given
in ref. 4. We emphasize that our value is consistent with exten-
sive observations of interconversion reactions involving plum-
bonacrite.
= -(1593.8 - 10.28 log pi - 1.14 log pi)
If) Calculation of solubility-stability relationships
Solubility calculations were performed using the computer
program DIST, which was developed at these laboratories by
Robert J. Lemire and Nava C. Garisto.' This program calcu-
= -(583.7 - 3.43 log pi - 0.38 log pi - 1.90 log pi) lates species distributions, at given pH and ligand activities,
from Gibbs energies of formation of the solid and solution
species concerned. The results are expressed as activities, in
(Estimated errors include our own experimental uncertainties, molal units. We have used molar units throughout the dis-
plus those in Gibbs energies of formation of the parent oxides, cussion, since the resulting errors are negligible at the level of
as indicated in ref. 24, since the latter uncertainties are not precision of our data. Our selected Gibbs energy data are
stated in ref. 21. Data from ref. 21 were used for consistency shown in Table 2.
with the solubility computations, shown below.)
Strictly speaking, this program calculates the activities of species
"The data of ref. 21 refer to a standard state pressure of one atmo- in saturated solutions. [t is reasonable to equate the total lead activity,
sphere (1.013 X 105 Pa). In eqs. [ I l l to [13], p',, p i , pi are partial {Pb,) = Cx{Pb,L,."), to solubility at ionic strengths below about 0.01
pressures in atm. mol dm-3.
TAYLOR AND LOPATA 399

Although hydroxo- and carbonato-lead(I1) complexes have Bilinski and Schindler (14) recently reported somewhat lower
received much experimental attention, some uncertainties formation constants (log p = 5.40 and 8.86, respectively, at
about them persist. The most recent detailed assessment of I = 0.3 mol dm-3).
thermochemical data for hydroxo-lead(I1) complexes appears It is convenient to discuss the solubility calculations in terms
to be that reported by Sadiq and Lindsay (21), and most of our of PC, and pPb,, the negative logarithms of {C,) and {Pb,),
data are from this source. These data originate mainly from the respectively (cf. pH). We computed species distributions at pH
work of Olin and Carell (27-30), as previously compiled by intervals of 0.5 and PC, intervals of 1.0. A finer grid was used
Baes and Mesmer (31) and by Smith and Martell (32). Sadiq to elucidate details of complex behaviour in some portions of
and Lindsay have adjusted the Gibbs energy data tabulated by the pH-PC, matrix. Results are summarized in Figs. 3 to 5.
Smith and Martell for internal consistency with their re- Figure 3 is a composite solubility diagram, showing pPb, and
calculated values for PbO(s) and Pb2+(aq). the stability fields of litharge and the three carbonates as func-
We have omitted two species which were included in Sadiq tions of pH and PC,. Figure 4 shows the computed fields of
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

and Lindsay's compilation: P ~ ( O H ) ; - ( ~and


~ ) Pb(OH),(s). The preponderance of the various aqueous lead species in saturated
former had been included on the basis of a single report by solutions, superimposed on solubility contours corresponding
Aksel'md (33), cited by Karapet'yants and Karapet'yants (34). to Figure 3. Figure 5 shows the computed solubilities of the
We found that its inclusion led to unrealistically high computed four solids, including conditions of their metastable occur-
lead concentrations at high pH, exceeding 1 mol dm-, above rence, as a function of pH at PC, = 4. Table 3 lists computed
about pH = 13.5. The data selected by Sadiq and Lindsay equilibrium pPb, and PC, values, as a function of pH, for the
imply that Pb(OH),(s) is stable relative to litharge and water by three stable solid assemblages (litharge t plumbonacrite;
26.0 kJ mol-' at 25°C. This is based on two earlier thermo- plumbonacrite + hydrocemssite; hydrocemssite t cemssite).
chemical compilations, but the original source is obscure.
However, there is considerable doubt about the very existence (g) Description of solubility-stability relationships
of solid Pb(OH),, many earlier reports being referable to basic Figure 3 illustrates the development of the lead carbonate
lead nitrates (35-37). We noted above that the hydrated stability fields as total carbonate activity increases. At PC, >
lead(I1) oxide, Pb60,(OH),, may be marginally stable with 9.5 and pPb, > 0 , litharge is the sole stable solid, with a
Can. J. Chem. 1984.62:395-402.

respect to PbO(s) and H20(1) (1 1). However, we have not minimum computed solubility of pPb, = 4.6 at pH = 10.3.
observed it, except in some electrochemical experiments under (See top of Fig. 3; note that a portion of the PbO solubility
conditions of high super~aturation.~ We omitted massicot, curve at PC, = 10 is obscured in this view of the solubility
since the consensus is that it is metastable with respect to surface.) Most of the litharge solubility surface at pH < 9 is
litharge at 25°C (2 1, 24). probably not accessible in practice without precipitation of
Two further species, P~,(oH): and P ~ ~ ( o H ) have~ + , recent- basic lead salts. As PC, decreases (i.e. total carbonate activity
ly been reported (38-40). When we included P~~(oH):in our increases), solid lead carbonates progressively suppress lead
data set, it led to unrealistically high computed solubilities for solubility, and their stability fields encroach into that of li-
PbO around pH = 9 to 11, with a minimum solubility of about tharge. Plumbonacrite initially appears at PC, = 9.5, hydro-
5 X lo-' mol dm-3. The computed minimum solubility when cemssite at PC, = 9.0, and cerussite at PC, = 5.5. The stability
P~,(oH); was omitted was about 2.5 x lo-' mol dm-,, in field boundaries (dashed lines in Fig. 3) progress to higher pH
better agreement with the experimental minimum solubility of as PC, decreases. Encroachment by the stability fields of the
litharge, 6.7 X lo-' mol dm-, (41). The species P~,(OH); may complex salts MOH-2PbC03 (not shown in Fig. 3) occurs at
thus be present, but its reported stability constant appears to be pH .= 10 and PC, < 1 (M = Na) or PC, < 0 (M = K), when
too high. We therefore omitted it from our final computations. Na and K are the counterions in the added carbonate salts.
We also omitted P~,(oH):+, since its formation constant has The preponderance diagram (Fig. 4) shows that polymeric
only been determined at high ionic strength (3.0 mol dm-, hydroxo-lead complexes are important only at about PC, > 8
LiClO,) (39, 40). The quality of agreement between Olin's and pH < 10 for realistic lead concentrations. At lower PC,
speciation model and experimental data indicates that the inclu-
sion of further species will only have minor effects on com-
puted solubility behaviour (27-30, 38-40). Ohtaki and co-
values, polymerization is suppressed by precipitation of the
solid carbonates. Lead solubility reaches a minimum, pPb, -
6.6, in slightly alkaline solution near PC, = 4; at still lower PC,
workers (39, 40) present evidence that the species formerly values solubility is enhanced by carbonate-lead(I1) complex
identified as P ~ ~ ( o H )is~in+ fact P~~o(oH):+,which has also formation. This behaviour is consistent with the solubility data
been observed in the solid state (10,42,43). We have incorpo- of Bilinski and co-workers (14, 46). The field of predominance
rated this suggestion by subtracting the Gibbs energy of for- of Pb(~0,):- (Fig. 4) indicates that the best area for further
mation of H20(1) from that recommended for P~~(oH):+in ref. experimental evaluation of the formation constant of this spe-
21, the two species being related by reaction [14]: cies would be near PC, = 1-2 and pH = 9- 1 1. The low and
relatively constant computed lead solubility within the field of
predominance of P ~ C indicates
O~ that solubility measurements
Several workers have measured formation constants, p, for are unlikely to shed further light on this complex.
the two lead carbonate complexes P ~ C Oand ~ P~(co,):-. There is some interest in whether cemssite or hydrocemssite
Clever and Johnston (44) tentatively recommended the re- is the stable form of lead carbonate in natural surface waters
spective values p = 2 x lo6 and 6 x lo9 at I = 0.1 mol dm-3, (14, 15). Our data indicate that cerussite should be the stable
as reported by Bilinski et al. (45). We have used these data, phase at the usual atmospheric p(C02)of = 30 Pa, but that only
corrected to zero ionic strength using eq. [8]. We note that slight depletion of C 0 2 would be necessary to enter the hydro-
cerussite stability field. However, Fig. 5 shows that the solu-
Unpublished work, in collaboration with D. W. Shoesmith and M. bilities are similar, especially near neutral pH. This similarity
G. Bailey. also explains the substantial uncertainty in p , , as determined
CAN. J . CHEM. VOL. 62, 1984
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.
Can. J. Chem. 1984.62:395-402.

FIG. 3. Computed solid/solution equilibrium surface in the lead spectively). Dashed lines represent stability boundaries between solid
carbonate system at 2S°C, expressed as negative logarithms of proton, phases in the sequence (left to right) cerussite, hydrocerussite, plum-
total lead and total carbonate activities (pH, pPb,, and PC,, re- bonacrite, litharge.

from solubility measurements (14), compared with direct ob-


servation of phase transformations. Interestingly, the minimum
computed lead solubility varies only within the narrow limits,
5.8 < pPb, < 6.6, as p(C0,) varies between lo-) and lo5 Pa,
although the position of the minimum progresses from pH = 10
to 6.5.
Most of our lead solubility measurements were in the range
pH > 10 and PC, < 3, and thus lay within the predominance
fields of P~(co,):- and Pb(0H);. These measurements were
not sufficiently precise to permit any further critique of the
computed solubilities, but the general trends paralleled the
behaviour described in the appropriate portions of Figs. 3
and 4.
(h) Phase equilibria above 25°C
We briefly investigated the solid phase interconversions at
50, 75, and 100°C, in solutions with [C,] near 0.1 mol dm-,.
We found only a slight temperature dependence for all three FIG. 4. Computed fields of preponderance of the various dissolved
equilibria, when expressed as concentrations of carbonate and lead species (bold outlines) in saturated solutions corresponding to the
bicarbonate or hydroxide. However, applying Helgeson's data equilibrium surface in Fig. 3. Numbered contours represent total lead
for the dissociation equilibria of water and carbonic acid at activity, expressed as pPb,.
TAYLOR AND LOPATA 40 1

By analogy with the computed high-temperature Gibbs ener-


gy data reported (24) for aragonite, strontianite, and witherite,
which are all isostructural with cerussite, and display similar
AfGO-T relationships, we would anticipate A,GO (cerussite, s ,
373 K) = -608 kJ mol-I, in reasonable agreement with our
experimental estimate. The temperature dependence of p, ap-
pears to differ substantially from that of p , and p,. Whereas p,
increases by only about an order of magnitude between 25 and
100°C, p, and p, both increase by about four orders of mag-
nitude. This is probably because H 2 0 is a reactant in the for-
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

ward reaction of equilibria [3] and [4], but a product in reaction


[5]. These p(C0,)-T relationships indicate that the stability
field of cerussite encroaches on that of hydrocerussite with
increasing temperature. The stability fields of hydrocerussite
FIG.5. Computed equilibrium dissolved lead activities for the four
solids litharge ( A ) , plumbonacrite (B), hydrocerussite (C), and ce- and plumbonacrite in the presence of H,0(1) are likely to disap-
russite (D) at PC, = 4, as a function of pH. pear at about 150 and 200°C, respectively. However, the phase
diagram will probably be further complicated in this region by
TABLE3. Computed ionic activities of saturated solutions in equi- the emergence of stability fields for the anhydrous basic car-
librium with lead oxide/carbonate phase assemblages bonates, nPbC0,. bPbO (8).

Litharge + Plumbonacrite + Hydrocerussite + Acknowledgements


plumbonacrite hydrocerussite cerussite
Part of this work was jointly funded by AECL and Ontario
Hydro under the CANDEV agreement. W e are grateful to
Can. J. Chem. 1984.62:395-402.

pH PC, pPb, PC, pPb, PC, pPb,


A. M. Duclos for performing many of the XRD measurements,
and to R. J . Lemire and J. Paquette for many helpful dis-
cussions.

1. E. W. ABEL.In Comprehensive inorganic chemistry. Vol. 2.


Edited by J. C. Bailar, H. J. Emeltus, R. Nyholm, and A. F.
Trotman-Dickinson. Pergamon Press. 1973. pp. 119- 123.
2. J. K. OLBY.J. lnorg. Nucl. Chem. 28, 2507 (1966).
3. R. M. CARRELS and C. L. CHRIST. In Solutions, minerals and
equilibria. Harper and Row, New York. 1965. pp. 233-238.
4. D. F. HAACKE and P. A. WILLIAMS. J. Inorg. Nucl. Chem. 43,
406 (1981).
5. J. YAMAGUCHI, Y. SAWADA, 0 . SAKURAI, K. UEMATSU, N.
M~ZUTANI, and M. KATO.Thermochim. Acta, 35, 307 (1980).
6. J. YAMAGUCHI, Y. SAWADA, 0 . SAKURAI, K. UEMATSU, N.
MIZUTANI, and M. KATO.Thermochim. Acta, 37, 79 (1980).
7. D. A. GRISAFE and W. B. WHITE.Am. Miner. 49, 1184 (1964).
8. G. TODDand E. PARRY. Nature, 202, 386 (1964).
9. R. A. H o w r ~and W. MOSER.Nature, 219, 372 (1968).
10. A. OLINand R. SODERQUIST. K. Tekn. Hogsk. Handl. No. 289,
517 (1972).
1 I. M. RANDALL and H. M. SPENCER. J. Am. Chem. Soc. 50, 1572
*Indicates impossibly high total lead or carbonate activity. (1928).
12. R. R. STASKO and G. A. VIVIAN. An overview of the Ontario
elevated temperatures (47) to the estimated equilibrium solu- Hydro Carbon-14 Control Program. 112 Proceedings of Inter-
tion concentrations, we obtain substantial increases in all three national Conference on Radioactive Waste Management,
equilibrium partial pressures, as expected. W e estimate the Winnipeg, September 1982. Canadian Nuclear Society, Toronto.
pp. 547-553.
following values at 100°C: 13. J. BOULTON (Editor). Management of radioactive fuel wastes: the
Canadian disposal program. Atomic Energy of Canada Ltd. Re-
port AECL-63 14. 1978.
14. H. BILINSKI and P. SCH~NDLER. Geochim. Cosmochim. Acta, 46,
921 (1982).
15. J. D. HEMand W. H. DURUM. J. Am. Water Works Assoc. 65,
Combining these estimates with high-temperature Gibbs ener- 562 (1973).
gy data for PbO, CO,, and H 2 0 from ref. 23, we obtain the 16. R. M. HARRISON and D. P. H. LAXEN.Chem. Brit. 16, 316
(1 980).
following estimates of Gibbs energies of formation: 17. R. M. CARRELS. Am. Miner. 42, 780 (1957).
AfGO(plumbonacrite, s , 373 K) = -5 120 * 50 kJ mol-' 18. A. W. MANNand R. L. DEUTSCHER.
(1980).
Chem. Geol. 29, 293

AfGO(hydrocerussite, s , 373 K) = - 1643 * 18 kJ mol-I 19. M. H. BROOKER, S. SUNDER, P. TAYLOR, and V. J . LOPATA.
402 CAN. I. CHEM. VOL. 62, 1984

Can. J. Chem. 61, 494 (1983). 34. M. KH. KARAPET'YANTS and M. L. KARAPET'YANTS. Thermo-
20. M. W. CHASE,J. L. CORNUTT, A. T. Hu, H. PROPHET,A. N. dynamic constants of inorganic and organic compounds. Ann
SYVERUD, and L. C. WALKER. J. Phys. Chem. Ref. Data, 3,311 Arbor-Humphrey, Ann Arbor. 1970.
(1974). 35. A. E. NEWKIRK and V. B. HUGHES.Inorg. Chem. 9,401 (1970).
21. M. SADIQand W. L. LINDSAY. Selection of standard free energies 36. W. KWESTROO, C. LANGEREIS, and H. A. M. VANHAL.J. Inorg.
of forination for use In soil chemistry. Tech. Bull. 134, Colorado Nucl. Chem. 29, 33 (1967).
State Univ. Expt. Stn., 1979. 37. J. ROBINand A. THEOLIER. Bull. Soc. Chim. Fr. 680 (1956).
22. G. W. BIRDand V. J. LOPATA.In Scientific basis for nuclear 38. R. N. SYLVAand P. L. BROWN.J. Chem. Soc. Dalton, 1577
waste management. Vol. 2. Edlted by C. J. M. Northrup, Jr. (1980).
Plenum, New York. 1980. pp. 419-426. 39. T. KAWAI,S. ISHIGURO, and H. OHTAKI. Bull. Chem. Soc. Jpn.
23. W. STUMM and J. J. MORGAN. Aquatic chemistry. 2nd ed Wlley, 53, 2221 (1980).
New York. 1981. 40. S. ISHIGURO and H. OHTAKI.Bull. Chem. Soc. Jpn. 54, 335
24. R. A. ROBIE,B. S. HEMINGWAY, and J. R. FISHER.U.S. Geol. (1981).
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

Surv. Bull. 1452 (reprinted with corrections, 1979). 41. A. B. GARRETT,S. VELLENGA, and C. M. FONTANA. J. Am.
25. C. W. DAVIES. In Ion association. Butterworths, London. 1962. Chem. Soc. 61, 367 (1939).
Chapt. 3. 42. T. G. SPIRO,D. H. TEMPLETON, and A. ZALKIN.Inorg. Chem.
26. L. H. ADAMIand K. C. CONWAY. U.S. Bureau of Mines Report 8, 856 (1969).
Inv. 6822, 7 pp (1966). 43. A. OLIN and R. SODERQUIST. Acta Chem. Scand. 26, 3505
27. A. OLIN.Acta Chem. Scand. 14, 126 (1960). (1972).
28. A. OLIN.Acta Chem. Scand. 14, 814 (1960). 44. H. L. CLEVER and F. J. JOHNSTON. J. Phys. Chem. Ref. Data, 9,
29 B. CARELLand A. OLIN.Acta Chem. Scand. 14, 1999 (1960). 751 (1980).
30. B. CARELL and A. OLIN.Acta Chem. Scand. 16, 2350 (1962). 45. H. BILINSKI,R. HUSTON,and W. STUMM. Anal. Chim. Acta, 84,
31. C. F. BAES,JR. and R. E. MESMER. In The hydrolysis of cations. 157 (1976).
Wiley-Interscience, New York. 1976. pp. 358-365. 46. H. BILINSKIand M. MARKOVIC. Croat. Chem. Acta, 50, 125
32. R. M. SMITHand A. E. MARTELL.Critical stability constants. (1977).
Vol. 4. Critical complexes. Plenum, New York. 1976. 47. H. C. HELGESON. J. Phys. Chem. 71, 3121 (1967).
33. N. V. AKSEL'RUD. Dokl. Akad. Nauk SSSR, 132, 1967 (1960).
Can. J. Chem. 1984.62:395-402.
This article has been cited by:

1. Pascal E. Reiller, Michaël Descostes. 2020. Development and application of the thermodynamic database PRODATA dedicated
to the monitoring of mining activities from exploration to remediation. Chemosphere 251, 126301. [Crossref]
2. Xiao Ma, Barbara H. Berrie. 2020. Lead Chlorides in Paint on a Della Robbia Terracotta Sculpture. Analytical Chemistry .
[Crossref]
3. Amin Deng, Jun Yuan, Chenze Qi, Yong Gao. 2020. Hollow SiO2 microspheres with thiol-rich surfaces: The scalable templated
fabrication and their application for toxic metal ions adsorption. Materials Chemistry and Physics 243, 122625. [Crossref]
4. Eva Mariasole Angelin, Sara Babo, Joana Lia Ferreira, Maria João Melo. 2019. Raman microscopy for the identification of
pearlescent pigments in acrylic works of art. Journal of Raman Spectroscopy 50:2, 232-241. [Crossref]
5. Luca Nodari, Laura Tresin, Alvise Benedetti, Maria Katia Tufano, Patrizia Tomasin. 2019. Conservation of contemporary art:
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

Alteration phenomena in a XXI century artwork. From contactless in situ investigations to laboratory accelerated ageing tests.
Journal of Cultural Heritage 35, 288-296. [Crossref]
6. Zhiyong Zhang, Chang Liu, John T. Brosnahan, Hua Zhou, Wenqian Xu, Sen Zhang. 2019. Revealing structural evolution of
PbS nanocrystal catalysts in electrochemical CO 2 reduction using in situ synchrotron radiation X-ray diffraction. Journal of
Materials Chemistry A 43. . [Crossref]
7. Yongliang Xiong, Leslie Kirkes, Terry Westfall, Jandi Knox, Cassandra Marrs, Heather Burton. 2018. A Pitzer Model for Lead
Oxide Solubilities in the Presence of Borate to High Ionic Strength. Journal of Solution Chemistry 47:12, 1905-1925. [Crossref]
8. Damon K. Roth, Jacob R. Wagner, David A. Cornwell. 2018. Impacts of Source Water Blending on Lead Release. Journal -
American Water Works Association 110:10, 15-25. [Crossref]
9. Oleg Siidra, Diana Nekrasova, Wulf Depmeier, Nikita Chukanov, Anatoly Zaitsev, Rick Turner. 2018. Hydrocerussite-related
Can. J. Chem. 1984.62:395-402.

minerals and materials: structural principles, chemical variations and infrared spectroscopy. Acta Crystallographica Section B
Structural Science, Crystal Engineering and Materials 74:2, 182-195. [Crossref]
10. Teresa Palomar, Emilio Cano. 2018. Comparative assessment of mechanical, chemical and electrochemical procedures for
conservation of historical lead. Journal of Cultural Heritage 30, 34-44. [Crossref]
11. Arturo Mendoza-Flores, Mario Villalobos, Teresa Pi-Puig, Nadia Valentina Martínez-Villegas. 2017. Revised aqueous solubility
product constants and a simple laboratory synthesis of the Pb(II) hydroxycarbonates: Plumbonacrite and hydrocerussite.
GEOCHEMICAL JOURNAL 51:4, 315-328. [Crossref]
12. Sheldon Masters, Gregory J. Welter, Marc Edwards. 2016. Seasonal Variations in Lead Release to Potable Water. Environmental
Science & Technology 50:10, 5269-5277. [Crossref]
13. Yongliang Xiong. 2015. Experimental determination of lead carbonate solubility at high ionic strengths: a Pitzer model
description. Monatshefte für Chemie - Chemical Monthly 146:9, 1433-1443. [Crossref]
14. Frederik Vanmeert, Geert Van der Snickt, Koen Janssens. 2015. Plumbonacrite Identified by X-ray Powder Diffraction
Tomography as a Missing Link during Degradation of Red Lead in a Van Gogh Painting. Angewandte Chemie International
Edition 54:12, 3607-3610. [Crossref]
15. Frederik Vanmeert, Geert Van der Snickt, Koen Janssens. 2015. Plumbonacrite Identified by X-ray Powder Diffraction
Tomography as a Missing Link during Degradation of Red Lead in a Van Gogh Painting. Angewandte Chemie 127:12, 3678-3681.
[Crossref]
16. A. G. Morachevskii. 2014. Physicochemical studies of utilization of lead batteries. Russian Journal of Applied Chemistry 87:3,
241-257. [Crossref]
17. Yin Wang, He Jing, Vrajesh Mehta, Gregory J. Welter, Daniel E. Giammar. 2012. Impact of galvanic corrosion on lead release
from aged lead service lines. Water Research 46:16, 5049-5060. [Crossref]
18. Salla H. Venäläinen. 2012. Sorption of lead by phlogopite-rich mine tailings. Applied Geochemistry 27:8, 1593-1599. [Crossref]
19. G. Falgayrac, S. Sobanska, C. Brémard. 2012. Particle–Particle Chemistry between Micrometer-Sized PbSO 4 and CaCO 3
Particles in Turbulent Flow Initiated by Liquid Water. The Journal of Physical Chemistry A 116:27, 7386-7396. [Crossref]
20. Salla H. Venäläinen. 2011. Apatite ore mine tailings as an amendment for remediation of a lead-contaminated shooting range
soil. Science of The Total Environment 409:21, 4628-4634. [Crossref]
21. J.H. Kyle, P.L. Breuer, K.G. Bunney, R. Pleysier, P.M. May. 2011. Review of trace toxic elements (Pb, Cd, Hg, As, Sb, Bi,
Se, Te) and their deportment in gold processing. Part 1: Mineralogy, aqueous chemistry and toxicity. Hydrometallurgy 107:3-4,
91-100. [Crossref]
22. Vesna Matović, Nada Vasković, Suzana Erić, Danica Srećković-Batoćanin. 2010. Interaction between binding materials—the cause
of damage to gabbro stone on the monument to the unknown soldier (Serbia). Environmental Earth Sciences 60:6, 1153-1164.
[Crossref]
23. Guillaume Falgayrac, Sobanska Sobanska, Jacky Laureyns, Claude Brémard. 2006. Heterogeneous chemistry between PbSO4
and calcite microparticles using Raman microimaging. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 64:5,
1095-1101. [Crossref]
24. Mohammed A Amin, Sayed S Abdel Rehim. 2004. Pitting corrosion of lead in sodium carbonate solutions containing NO3−
ions. Electrochimica Acta 49:15, 2415-2424. [Crossref]
25. M. E. Essington, J. E. Foss, Y. Roh. 2004. The Soil Mineralogy of Lead at Horace's Villa. Soil Science Society of America Journal
68:3, 979-993. [Crossref]
26. Jean Tétreault, Emilio Cano, Maarten van Bommel, David Scott, Megan Dennis, Marie-Geneviève Barthés-Labrousse, Léa
Downloaded from www.nrcresearchpress.com by 157.41.163.159 on 03/17/20. For personal use only.

Minel, Luc Robbiola. 2003. Corrosion of Copper and Lead by Formaldehyde, Formic and Acetic Acid Vapours. Studies in
Conservation 48:4, 237-250. [Crossref]
27. Geoffrey C. Allen, Leon Black, Philip D. Forshaw, Nigel J. Seeley. 2003. Raindrops Keep Falling on My Lead. Journal of
Architectural Conservation 9:1, 23-44. [Crossref]
28. Sang-Yul Kim, Nobutoshi Tanaka, Toshihiko Matsuto. 2002. Solubility and adsorption characteristics of Pb in leachate from
MSW incinerator bottom ash. Waste Management & Research 20:4, 373-381. [Crossref]
29. Mark Pritzker. 2000. Pore transport-controlled shrinking-core systems involving diffusion, migration, and homogeneous
reactions: Part II. Application of model for PbSO4-carbonate system to experimental data. Metallurgical and Materials
Transactions B 31:4, 693-703. [Crossref]
30. Jean Tétreault, Jane Sirois, Eugénie Stamatopoulou. 1998. Studies of lead corrosion in acetic acid environments. Studies in
Can. J. Chem. 1984.62:395-402.

Conservation 43:1, 17-32. [Crossref]


31. Susan M. Grimes, Simon R. Johnston, David N. Batchelder. 1995. Lead carbonate–phosphate system: solid–dilute solution
exchange reactions in aqueous systems. The Analyst 120:11, 2741-2746. [Crossref]
32. Y. Gong, J.E. Dutrizac, T.T. Chen. 1992. The conversion of lead sulphate to lead carbonate in sodium carbonate media.
Hydrometallurgy 28:3, 399-421. [Crossref]
33. R.g. Barradas, T.J. Vandernoot. 1987. Two-dimensional electrocrystallisation oif PbCO3 and Pb3(CO3)2(OH)2 on lead amalgam
electrodes in sodium carbonate and bicarbonate solutions. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry
233:1-2, 237-250. [Crossref]
34. M. Tomlinson. 1985. Solubility phenomena in industrial and natural system. Journal of Solution Chemistry 14:7, 443-456.
[Crossref]
35. Nona J. Flemming, Vincent J. Lopata, Barbara L. Sanipelli, Peter Taylor. 1984. Thermal decomposition of basic lead carbonates:
A comparison of hydrocerussite and plumbonacrite. Thermochimica Acta 81, 1-8. [Crossref]

You might also like