08.02.2015 Views

Homogeneous precipitation from solution by urea hydrolysis: a ...

Homogeneous precipitation from solution by urea hydrolysis: a ...

Homogeneous precipitation from solution by urea hydrolysis: a ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Homogeneous</strong> <strong>precipitation</strong> <strong>from</strong> <strong>solution</strong> <strong>by</strong> <strong>urea</strong> <strong>hydrolysis</strong>: a novel chemical route<br />

to the a-hydroxides of nickel and cobalt<br />

Mridula Dixit," Gonur N. Subbanna' and P. Vishnu Kamath""<br />

"Department of Chemistry, Central College, Bangalore University, Bangalore 560 001, India<br />

bMaterials Research Centre, Indian Institute of Science, Bangalore 560 01 2, India<br />

<strong>Homogeneous</strong> <strong>precipitation</strong> <strong>from</strong> <strong>solution</strong> <strong>by</strong> <strong>hydrolysis</strong> of <strong>urea</strong> at elevated temperatures (T= 120 "C) yields novel ammoniaintercalated<br />

a-type hydroxide phases of the formula M(OH), (NH,),,, (H,O),( N003), -x where x = 2, y = 0.68 for M = Ni and<br />

x = 1.85, y = 0 for M = Co. These triple-layered hexagonal phases (a = 3.08 & 0.01 A, c = 21.7 & 0.05 A) are more crystalline than<br />

similar phases obtained <strong>by</strong> chemical <strong>precipitation</strong> or electrosynthesis. This method can be adapted as a convenient chemical route<br />

to the bulk synthesis of a-hydroxides.<br />

The hydroxides of nickel and cobalt find applications as<br />

electrode materials for alkaline secondary These<br />

hydroxides have a hexagonal layered structure and exist in<br />

two polymorphic forms, a and p.3 The a-hydroxifles of nickel<br />

and cobalt have a. large interlayer spacing (> 7 A) compared<br />

to the p form (4.6 A) and are known to exhibit a higher charge<br />

capacity and better discharge efficiency. However, the a-<br />

hydroxides are metastable and are difficult to synthesize as<br />

they age rapidly to the p form during synthesis or on storage<br />

in strong alkali. LeBihan et uL4 reported the first chemical<br />

synthesis of a-nickel hydroxide <strong>by</strong> <strong>precipitation</strong> with liquid<br />

NH, followed <strong>by</strong> rapid centrifugation. Genin et aL5 followed<br />

a similar procedure, but in both studies, the material produced<br />

was poorly ordered and displayed broad features in the X-ray<br />

diffraction patterns. The most reliable synthetic route to a-<br />

nickel hydroxide is the electrosynthetic route <strong>by</strong> the cathodic<br />

reduction of nickel nitrate <strong>solution</strong>s.6 However, this reaction<br />

is current density dependent and the a-hydroxide forms only<br />

at low current densities, there<strong>by</strong> making electrosynthesis a<br />

very slow and inappropriate route to the bulk synthesis of a-<br />

nickel hydroxide. There has been, therefore, a considerable<br />

degree of interest in evolving an accelerated chemical route to<br />

bulk a-hydroxides of nickel and cobalt. Organic additives such<br />

as glucose, fructose, lactose and citric acid were found to<br />

mediate the <strong>precipitation</strong> of the a-hydroxides of nickel and<br />

cobalt.',' However, the products were found to have very poor<br />

crystallinity.<br />

Of late, homogeneous <strong>precipitation</strong> <strong>from</strong> <strong>solution</strong> <strong>by</strong><br />

<strong>hydrolysis</strong> of <strong>urea</strong> has been increasingly employed to synthesize<br />

novel hydroxide phases' and fine-particle<br />

There<br />

are two reports in the literature of the use of this method for<br />

the synthesis of nickel hydroxide. Maruthiprasad et report<br />

the synthesis of nickel trihydroxyisocyanate, while Avena<br />

et characterize their product as a Pb,-type nickel hydroxide,<br />

where Pbc stands for a poorly crystallised p phase. While the<br />

diffraction and spectral data presented <strong>by</strong> both groups are<br />

similar, the inconsistency in their conclusions prompted us to<br />

reinvestigate this simple method of preparation, especially as<br />

the published data appear to point towards the formation of<br />

a-type hydroxides.<br />

Our investigations reveal that the product is a novel NH,<br />

intercalated a-type hydroxide. We have extended this method<br />

to the synthesis of a-cobalt hydroxide, a phase first synthesized<br />

in our laboratory.' In addition, <strong>urea</strong> <strong>hydrolysis</strong> at elevated<br />

temperatures (12O"C), as we have achieved, yields a highly<br />

ordered product with very low reaction times (100 min). We<br />

account for all the observed structural, compositional and<br />

morphological features of the a-type hydroxides.<br />

Experimental<br />

Synthesis<br />

To 50ml of the metal nitrate <strong>solution</strong> (0.5mol dm-3) was<br />

added 12 g of <strong>urea</strong>. <strong>Homogeneous</strong> <strong>precipitation</strong> was affected<br />

<strong>by</strong> placing this <strong>solution</strong> in a glass container in a stainless-steel<br />

domestic pressure cooker (100 min, 120 "C, 1 kg ern-,). The<br />

resulting precipitates were filtered, washed free of unreacted<br />

<strong>urea</strong> and dried to constant mass at 65°C.<br />

Wet chemical analysis<br />

The metal content of the hydroxide samples was estimated<br />

gravimetrically. The total base content was estimated <strong>by</strong><br />

dissolving a known quantity of the sample in an excess of<br />

hydrochloric acid (0.4 mol drn-,) and back-titrating the excess<br />

acid against standard sodium hydroxide using a pH meter.<br />

Intercalated NH3 was estimated <strong>by</strong> the microkjeldah procedure.<br />

A known quantity of the sample was suspended in<br />

10 ml of 40 massoh sodium hydroxide <strong>solution</strong>. Ammonia was<br />

liberated <strong>by</strong> steam distillation and absorbed in an excess of<br />

2% boric acid <strong>solution</strong>. This was titrated against standard<br />

sulfuric acid (0.025 mol drn-,) using a mixed (1 : 3) methylene<br />

blue-methylene red indicator. The microkjeldah procedure was<br />

standardized using a standard ammonium sulfate <strong>solution</strong>. All<br />

chemical titrations were performed at least three times to get<br />

concordant results.<br />

The difference between the total base and the intercalated<br />

NH, yields the hydroxy ion content of the sample. The net<br />

difference in the positive (Ni2+) and negative charges (OH-)<br />

was compensated <strong>by</strong> the inclusion of nitrate ions and finally<br />

the unaccounted mass was assigned to the intercalated water<br />

con tent.<br />

Characterisation<br />

Powder X-ray diffraction patterns (XRD) were recor9ed on a<br />

Philips X-rayo diffractometer using Cu-Ka (A = 1.541 A) or Co-<br />

Ka (1 = 1.79 A) radiation. The X-ray data were not of a quality<br />

that permitted least-squares refinement of peak positions and<br />

the patterns were indexed <strong>by</strong> a trial and error method starting<br />

with the lattice parameters of the model compounds reported<br />

previ~usly.'~ IR spectra were recorded using a Nicolet Impact<br />

400D FTIR spectrometer using the KBr pellet method at a<br />

re<strong>solution</strong> of 3 cm- '. Thermogravimetric studies were carried<br />

out on a home-made system at a heating rate of 2.5 "C min-'.<br />

Isothermal losses were recorded after holding the sample at<br />

the desired temperature for 2 h. Electron microscopy was<br />

J. Muter. Chem., 1996, 6(8), 1429-1432 1429


~~ ~<br />

carried out using a JEOL 200 CX electron microscope Finely<br />

powdered samples were dispersed onto carbon mesh (size 200)<br />

for microscopic investigations<br />

Results<br />

In Fig 1A and B are shown powder XRD patterns of the<br />

samples obtained <strong>by</strong> homogeneous <strong>precipitation</strong> <strong>from</strong> nickel<br />

nitrate and cobalt nitrate <strong>solution</strong>s, respectively The prominent<br />

d spacings are listed in Table 1 together with those of a-nickel<br />

hydroxide obtained <strong>from</strong> the literature It is clear that the<br />

hydroxides obtained <strong>from</strong> homogeneous <strong>precipitation</strong> are a-<br />

type phases which c!n be indexed on triple-layered hexagonal<br />

cell (a= 3 08 +O 01 A, c=21 7_+005 A) In contrast, the more<br />

stable p-hydroxides obtained <strong>from</strong> conventionF1 chemica! synthesis<br />

have a much smaller unit cell (a= 3 10 A, c=4 76 A)<br />

As there has been considerable discussion in the literature<br />

on the presence of intercalated anions in the a-type hydrox-<br />

ides16 and especially among the products of <strong>urea</strong> <strong>hydrolysis</strong><br />

reactions,13 l4 a detailed chemical analysis of the samples was<br />

carried out, the results of which are summarised in Table2<br />

Estimation of the metal content clearly showed that the<br />

samples were anything but stoichiometric divalent hydroxides<br />

The observed metal content was much lower than what would<br />

be expected of a stoichiometric divalent hydroxide (63 3%),<br />

throwing open the possibility of the presence of several intercalated<br />

species The a-hydroxides generally contained significant<br />

amounts of intercalated anions with a matching deficiency in<br />

the hydroxy ion contentI6 The total base contents of the<br />

samples were therefore estimated using chemical titrations<br />

The results were quite surprising The hydroxides of both<br />

cobalt and nickel revealed a total base content in excess of<br />

what would be expected <strong>from</strong> the metal content As the excess<br />

base has to be of the neutral variety, the presence of intercalated<br />

NH,, a product of <strong>urea</strong> <strong>hydrolysis</strong>, was surmised Accordingly,<br />

microkjeldah estimations of ammoniacal nitrogen revealed the<br />

presence of up to 04 moles of NH, in both the hydroxides<br />

The difference in the chemical and microkjeldah estimations,<br />

which corresponds to the hydroxy ion content, was found to<br />

be stoichiometric for nickel hydroxide and deficient for cobalt<br />

Table2 Results of wet chemical analysis of the hydroxides obtained<br />

<strong>by</strong> homogeneous <strong>precipitation</strong> (hp)<br />

hp nickel<br />

hydroxide<br />

hp cobalt<br />

hydroxide<br />

e[ I I I<br />

lu<br />

Y<br />

15 35 55<br />

metal content (%) 51 9 55 1<br />

total base (Yo) 36 9 36 3<br />

NH, content (YO) 64 68<br />

net hydroxy content (YO) 30 6 29 6<br />

anion content (YO) 0 87<br />

water content (%) 109 0<br />

net mass loss (%) 32 8 (33 2) 24 8 (24 9)<br />

Values in parentheses are calculated on the basis of the proposed<br />

formula (see text)<br />

Table 3 Isothermal mass losses observed at different temperatures for<br />

the products of homogeneous <strong>precipitation</strong><br />

hp nickel hydroxide<br />

hp cobalt hydroxide<br />

I I I I I 1<br />

15 35 55<br />

2Bldegrees<br />

Fig. 1 A, Powder X-ray diffraction patterns of homogeneously precipitated<br />

nickel hydroxide (a) and nickel oxide, Ni0 (b) obtained <strong>by</strong><br />

thermal dccomposition of the hydroxide Data recorded using Cu-Ka<br />

(A= 1 541 A) radiation B, Powder X-ray diffraction patterns of<br />

homogeneously precipitated cobalt hydroxide (a) and cobalt oxide,<br />

Co,O, (b) obtained <strong>by</strong> thermal Gecomposition of the hydroxide Data<br />

recorded using Co-Ka (A = 1 79 A) radiation<br />

T/”C mass loss (YO) T/”C mass loss (YO)<br />

190<br />

230<br />

250<br />

270<br />

330<br />

370<br />

450<br />

36 150 36<br />

103 170 147<br />

20 7 190 20 6<br />

27 2 210 22 9<br />

30 0 310 24 8<br />

32 1<br />

32 8<br />

Table 1 X-ray powder diffraction data of the products of homogeneous<br />

<strong>precipitation</strong> (hp)<br />

loot<br />

I<br />

hp nickel hp cobalt<br />

hkP a-~l(~~),b hydroxide‘ hydroxided<br />

003 7 79 7 26 7 35<br />

006 3 908 3 616 3 612<br />

101 2 676 2 673 2 682<br />

012 2 604 - 2 598<br />

009 - 2 423 2 410<br />

015 2 321 - 2 282<br />

110 1541 1551 -<br />

“The observed patterns were indexed according tq these hkl planes %n<br />

a hexagonal cell bFrom ref 15 :a = 3 09 _+ 0 01 A, c = 21 77 f 0 05 A<br />

da=308+001 A, ~=2166+005A<br />

70t<br />

I I , I I<br />

0 100 200 300 400<br />

TIOC<br />

Fig. 2 TG data for homogeneously precipitated nickel hydroxide (a)<br />

and cobalt hydroxide (b)<br />

1430 J Muter Chem, 1996, 6(8), 1429-1432


hydroxide. The deficiency in the latter case was assumed to be<br />

made up <strong>by</strong> the presence of nitrate ions to achieve charge<br />

neutrality. The unaccounted-for mass was made up <strong>by</strong> the<br />

inclusion of water molecules. The analysis yields approximate<br />

molecular formulae Ni(OH)2(NH3)o 4(H20)0 68 and<br />

Co(OH), 85(NH3)o 4(N03)0 15 for the products obtained <strong>from</strong><br />

nickel and cobalt nitrate <strong>solution</strong>s, respectively.<br />

These formulae were verified <strong>by</strong> recording the isothermal<br />

mass losses of the samples at various temperatures (see Table 3).<br />

Both samples were decomposed completely <strong>by</strong> 400 "C, and the<br />

mass losses after decomposition matched those expected <strong>from</strong><br />

the above formulae (nickel sample: obs. 32.8%, exptd 33.2%;<br />

cobalt sample: obs. 24.8%, exptd. 24.9%). The expected mass<br />

losses were calculated based on the observation of the formation<br />

of Ni0 and Co304, respectively, as the products of<br />

thermal decomposition (see Fig. 1 for XRD data). However,<br />

these hydroxides showed a single-step dehydration and<br />

decomposition in their thermograms (see Fig. 2).<br />

In contrast, chemical analysis of a control sample of P-nickel<br />

hydroxide yielded the formula Ni(OH),, with a net isothermal<br />

mass loss of 18.1% (exptd. 19.4%), and a-nickel hydroxide<br />

yielded the formula Ni(OH)2(H,0)o,, with a total mass loss<br />

of 29% (exptd. 28.5%).<br />

In Fig. 3 the IR spectra of the samples obtained <strong>from</strong> nickel<br />

and cobalt nitrate <strong>solution</strong>s are shown. The spectra are typical<br />

of the a-type hydroxides which are described in detail elsewhere.16<br />

The only difference is the existence of an absorption<br />

at 2200 cm-l, which other authors13 have attributed to the<br />

existence of intercalated NCO- or CN- species in addition to<br />

the peaks due to intercalated nitrate anions seen in the<br />

1400-1000 cm-I region.<br />

In Fig. 4 the results of electron microscopic investigations<br />

1 I<br />

4000 3500 3000 2500 2000 1500 1000 500<br />

wavenumberkm<br />

Fig. 3 IR spectra of homogeneously precipitated nickel hydroxide (a)<br />

and cobalt hydroxide (b)<br />

on the samples obtained <strong>from</strong> nickel and cobalt nitrate <strong>solution</strong>s<br />

are presented. The nickel hydroxide sample shows a<br />

fibrillar turbostratic morphology [Fig. 4(u)]. The ring electron<br />

diffraction pattern taken <strong>from</strong> the same area at low magnification<br />

is shown in Fig. 4(b). However, the cobalt hydroxide<br />

particles are seen to have a needle-shaped morphology<br />

[Fig. 4(c)], and the corresponding electron diffraction pattern<br />

is sharp and spotty [Fig. 4(d)]. This is very much in contrast<br />

to the hexagonal platelet morphology seen among the p-<br />

Fig. 4 (a) Electron micrograph of homogeneously precipitated nickel hydroxide, (b) electron diffraction pattern of (a) (c) Electron micrograph of<br />

homogeneously precipitated cobalt hydroxide, (d) electron diffraction pattern of (c)<br />

J. Muter. Chern., 1996,6(8), 1429-1432 1431


hydroxides The turbostratic morphology is characteristic of<br />

the a-hydroxides<br />

Discussion<br />

Based on the observation of a strong absorption at 2200 cm-l<br />

in the IR spectra, Maruthiprasad et all3 characterised the<br />

product of homogeneous <strong>precipitation</strong> <strong>from</strong> nickel nitrate<br />

<strong>solution</strong> as nickel trihydroxyisocyanate [Ni2(0H),( NCO)]<br />

However, the IR spectra published <strong>by</strong> them contain equally<br />

strong evidence for the presence of intercalated nitrate ions, as<br />

do ours If intercalated anions are indeed present, based on<br />

charge neutrality considerations a matching deficiency in the<br />

hydroxy content can be expected However, chemical titrations<br />

reveal an excess base content Urea <strong>hydrolysis</strong> results in the<br />

formation of free NH,10-12 which under elevated pressures, as<br />

are obtained at 120 "C, can be expected to become intercalated<br />

within the hydroxide layers, leading to the ready formation of<br />

a-type phases The mole percentage of intercalated NH, is<br />

considerable, as revealed <strong>by</strong> the microkjeldah procedure While<br />

the cobalt hydroxide does show a 75% deficiency in the<br />

hydroxy content, the nickel hydroxide exhibits a stoichiometric<br />

hydroxy ion content It is our suggestion that the strong<br />

absorptions seen in the 1600-1000cm-1 region as well as at<br />

2200 cm-l arise <strong>from</strong> both intercalated and adsorbed species<br />

The X-ray diffraction data prove unequivocally that the<br />

phases obtained <strong>by</strong> homogeneous <strong>precipitation</strong> are a-type<br />

hydroxides The a-hydroxides are poorly ordered phases and<br />

exhibit broad bands in their X-ray diffraction patterns As a<br />

result, there is a considerable variation in the interlayer spacing<br />

reported <strong>by</strong> different authors3 The a phases obtained <strong>by</strong> us<br />

are rFasonably well ordered and the interlayer spacing observed<br />

(73 A) is less than the range of reported values l7 The<br />

presence of intercalated NH, in layered hydroxides is not<br />

unknown In the present instance, intercalated NH, appears<br />

to profoundly influence the thermal behaviour of the a-type<br />

hydroxides While a-hydroxides of nickel and cobalt are known<br />

to undergo multi-step mass losses,' l7 the samples obtained<br />

<strong>from</strong> homogeneous <strong>precipitation</strong> undergo a single-step dehydrationdecomposition<br />

reaction The observed mass loss agrees<br />

well with what is expected <strong>from</strong> the formulae obtained <strong>from</strong><br />

chemical analysis<br />

Electron microscopy results also show that the samples<br />

obtained <strong>from</strong> homogeneous <strong>precipitation</strong> exhibit a morphology<br />

more like the a-hydroxides, in contrast with the<br />

hexagonal platelet morphology of the 0-type hydroxides<br />

Conclusions<br />

We report here a simple chemical route to the bulk synthesis<br />

of a-hydroxides of nickel and cobalt <strong>by</strong> homogeneous <strong>precipitation</strong><br />

<strong>from</strong> <strong>solution</strong> <strong>by</strong> the <strong>hydrolysis</strong> of <strong>urea</strong> The phases<br />

which contain intercalated NH, are stable during synthesis<br />

and workup and this method can be adapted for the synthesis<br />

of nickel hydroxide positive electrodes for alkaline secondary<br />

battery applications<br />

Two of us (M D and P V K ) thank the Department of Science<br />

and Technology, Govt of India for financial support MD<br />

thanks the Council for Scientific and Industrial Research for<br />

the award of a Senior Research Fellowship<br />

References<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

K Watanabe, T Kikuoka and N Kumagai, J Appl Electrochem,<br />

1995,25,219<br />

J Bauer, D H Buss, H J Harms and 0 Glemser, J Electrochem<br />

SOC , 1990,137,173<br />

P Oliva, J Leonardi, J F Laurent, C Delmas, J J Braconnier,<br />

M Figlarz and F Fievet, J Power Sources, 1982,8,229<br />

S LeBihan, J Guenot and M Figlarz, C R Acad Sci Ser C, 1970,<br />

270,2131<br />

P Genin, A Delahaye-Vidal, F Portemer, K Tekia-Elhsissen and<br />

M Figlarz, Eur J Solid State Inorg Chem , 1991,28, 505<br />

K C Ho and J Jorne, J Electrochem SOC , 1990,137,149<br />

P V Kamath, J Ismail, M F Ahmed, G N Subbanna and<br />

J Gopalakrishnan, J Muter Chem , 1993,3,1285<br />

J Ismail, M F Ahmed, P V Kamath, G N Subbanna, S Uma<br />

and J Gopalaknshnan, J Solid State Chem , 1995, 114, 550<br />

R J Candal, A E Regazzoni and M A Blesa, J Muter Chem,<br />

1992,2,657<br />

J L Shi and J H Gao, J Muter Sci, 1995,30,793<br />

K Kandori, M Toshioka, H Nakashima and T Ishikawa,<br />

Langmuir, 1993,9, 1031<br />

K Kandori, H Nakashima and T Ishikawa, J Colloid Interface<br />

Sci , 1993,160,499<br />

B S Maruthiprasad, M N Sastri, S Rajagopal, K Seshan,<br />

K R Krishnamurthy and T S R P Rao, Proc Ind Acad Scz,<br />

1988,100,459<br />

M J Avena, M V Vazquez, R E Carbino, C P DePauli and<br />

V A Macagno, J Appl Electrochem, 1994,24,256<br />

J J Braconnier, C Delmas, C Fouassier, M Figlarz, B Beaudouin<br />

and P Hagenmuller, Rev Chzm Mzner , 1984,21,496<br />

F Portemer, A Delahaye-Vidal and M Figlarz, J Electrochem<br />

Soc , 1992,139,671<br />

B Mani and J P deNeufville, J Electrochem Soc , 1988,135,800<br />

P Benard, J P Auffredic and D Louer, Thermochim Acta, 1994,<br />

232,65<br />

Paper 6/03062I, Received 30th April, 1996<br />

1432 J Mater Chem , 1996, 6(8), 1429-1432

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!