Sie sind auf Seite 1von 23

ELS EVI ER Applied Clay Science 9 (1994) 165-187

Thermal stability and pozzolanic activity of calcined


kaolin
Changling He a'*, Emil Makovicky a, Bjarne Osb~eckb
aDepartment of Mineralogy, Institute of Geology, University of Copenhagen, Oster VoMgade 10,
Dk-1350 Copenhagen, Denmark
bTechnical Division, F.L. Smidth and Co. A/S Vigerslev AUe 77, DK-2500 Valby, Denmark
Received 14 October 1993; accepted after revision 26 April 1994

Abstract

Kaolins with fairly well crystallized kaolinite as the main component were calcined at respectively
550, 650, 800 and 950°C for 100 minutes. Both the raw and calcined samples, before and after being
mixed with either ordinary Portland cement or with Ca(OH)2 or after extended immersion in the
simulated cement paste pore solution saturated with Ca (OH)2, were studied by DTA (for raw kaolin),
XRD, SEM and EMPA. The compressive strength, chemical shrinkage and alkali and acid solubility
of samples were also investigated. The optimum calcination temperature from the combined scientific
and economic standpoint is 550°C, corresponding to the culminating point of kaolinite dehydroxy-
lation.

I. Introduction

Fired clays, ground and mixed with lime, formed the first man-made hydraulic cement
(Cook, 1985). Although encountering considerable competition from industrial waste prod-
ucts such as fly ash, fired clays are still a good choice as an energy saving admixture for
cement in many countries rich in clays but devoid of suitable waste material or natural
pozzolan.
Industrial grade clays or shales are usually a mixture of several clay minerals associated
with some non-clay minerals. However, it is useful to characterize the roles individual clay
species play in the mixture. This report represents the first part of the project "Pozzolanic
activity of calcined clays" that includes a series of rather pure clays: kaolinite, illite, Ca-
and Na-montmorillonite, sepiolite and mixed layer smectite-mica. All samples except

*Corresponding author.

0169-1317/94/$07.00 © 1994 Elsevier Science B.V. All rights reserved


SSDIO169-1317(94)OOOI3-G
166 ('. He et at. /Applied Clax Science 9 (I 994) 165-.187

kaolinite were provided as "'standard clays" by The American Clay Minerals Society. Two
fundamental approaches were chosen for the present study:
( 1) Laboratory investigations of chemical reactions between calcined clays and ordinaD'
Portland cement (OPC), laboratory reagent calcium hydroxide Ca(OH)2 (CH), or a solu-
tion that models the pore liquid of cement paste and saturated with CH (abbreviated hereafter
as SCPS).
(2) The investigation of pozzolanic activity, as described in ASTM (American Society
For Testing and Materials) Standards C618, C311 and C 109 and European Standard EN 196.
Various aspects of kaolin as a pozzolan have been studied (Mielenz et al., 1950; Forrester,
1975; Murat, 1983; Serry et al., 1984; Ayub et al., 1988; Gao and Zhang, 1989; De Silva
and Glasser, 1992; De Silva and Glasser, 1993). The present investigation, in which a
number of different methods have been applied to the same material, concentrates primarily
upon the connection between the mineralogy of the starting clay materials, the effects of
calcination on these materials and the properties of the newly produced phases obtained in
the pozzolanic reactions.

2. Materials and experimental methods

The kaolin sample used in the present study was provided by the Royal Porcelain Factory
of Denmark. This material was air dried and slightly pulverized prior to study. OPC provided
by F.L. Smidth Denmark A/S was used as reference and basic material for pozzolan
mixtures. Its chemistry, determined by X-ray fluorescence is presented in Table 1.
Bulk chemical analysis of the kaolin was carried out by X-ray fluorescence (multichannel
X-ray spectrometer, Philips PW 1606) and atomic absorption and potentiometric titration
(for Fe 2÷ ). Ignition loss was determined by heating to 980°C. Thermal analyses were run
on a differential thermal analyzer, Stanton Redcroft 673-4; with detectors for quantitative
determination of evolved CO2, H20 and SO2 and on a Mettler Recording Vacuum Ther-
moanalyzer with simultaneous recording of DTA, TG and DTG curves. A temperature
increase of 10°C per minute was used in both instruments. X-ray powder diffraction analyses
were performed on a Philips analytical powder diffractometer with an automatic divergence
slit, graphite monochromator and CuKa radiation. Non-oriented specimens were prepared
for general XRD but for the characterization of starting clays oriented specimens were also
prepared by sedimentation of the < 2/xm fraction on a glass slide. The quantitative min-
eralogical composition estimate for the starting kaolin was calculated by distributing the

Table 1
Chemistryand potentialmineralogyof ordinaryPortlandcement(wt%)

SiO2 A1203 Fe203 CaO MgO Mn203 TiO2 P205 K20 Na20

21.0 5.21 3.14 65.2 0.74 0.04 0.26 0.18 0.59 0.14

SrO SO3 LOI Total C3S C2S C3A C4AF CaSO4

0.14 2.45 0.75 99.84 59.0 16.0 8.5 9.6 4.2


C. He et al. /Applied Clay Science 9 (1994) 165-187 167

K20, A1203, SiO2 and H20 (released at about 550°C) contents from the bulk chemical
analysis and DTG among kaolinite, muscovite, microeline and quartz detected by XRD.
Calcination of samples was performed in a SCANDIA OVEN MC-80 equipped with a
microprocessor for automatic temperature control. The heating rate was about 7°C/min
over the first 100°C, but it decreased with increasing temperature. The average rate to 400°C
was about 5.7°C/min. and from 400°C to 800°C, about 1.3°C/min. The samples were
loosely spread on inert plates at a thickness of about 7 mm. After annealing at the desired
temperature for 100 minutes, samples were left to cool to room temperature in a closed
oven. The rate of cooling averaged about 10°C/h. To keep them protected from moisture,
samples were stored in double plastic bags as soon as they were cool enough to handle.
Calcined samples were ground in a SIEBTECHNIK swing mill for 30 seconds. The
specific surface of the samples was determined by nitrogen absorption (BET) and air
permeability (Blaine). The particle size distribution was determined by laser diffractometry
(Sympatec HELOS) after dispersion by ultrasonic treatment for 1 minute.
An attempt to pelletize calcined samples was unsuccessful due to the lack of cohesion.
The pellets used in the current experiments were made from the raw sample and calcined
subsequently. The composition and pH value of SCPS, in which the pellets were immersed,
simulated the pore liquid of cement paste when the ratio of water/cement is equal to 0.5
and after mixing for 24 hours. It contained 43mM Na and 95mM K, with a pH value equal
to 13.2. Additional CH was added to keep the solution saturated. The pellets were kept in
the solution for 7, 28 and 90 days respectively at 50°C before being analyzed.
The methods for measurement of alkali-soluble silica, used for chemical assessment of
the pozzolanic activity of fired clays, were described by Surana and Joshi (1990). Two
grams of powdered sample were boiled in 100 ml 0.5N Na(OH) for 3 minutes and the
amount of silica and alumina solubilized in the alkali solution measured by atomic absorption
spectroscopy (AAS). The measurement of acid-soluble silica and alumina was carried out
by soaking 2 grams of powdered sample in 100 ml of 24% HCI at 5°C and shaking for 10
minutes. After centrifugation, the supernatant was decanted for AAS.
Chemical shrinkage tests were carried out following Geiker and Knudsen (1982). The
samples were mixed with OPC and CH respectively. Two gram charges were made from
each specimen. The kaolin:OPC or kaolin:CH ratios were 1.0; the solution demand (same
solution as used for pellets described above) was 2 and 2.5 ml respectively for fair worka-
bility. The paste was manually mixed in a glass tube ( 17 mm inner diameter and 50 mm
height) and topped with paraffin oil. The open end was connected to a pipette with the scale
on which the oil level was read to monitor the changes in the specific volume of the mixture.
The tubes were immersed in a thermostatted water bath at 40°C for 45 days.
The compressive strength tests were a slightly modified version of those in ASTM C
109-87 and European Standard EN196. Twenty wt% of OPC was replaced by (meta) kaolin;
the ratio of sand to cement + pozzolan was 2.75 and the ratio of water to cement + pozzolan
varied between 0.52 and 0.6 for kaolin calcined at different temperatures in order to achieve
a similar degree of flow for all specimens. Mortars were cast as 5 cm cubes and were cured
in the moulds in a cabinet with 100% relative humidity at 23°C for the first day and after
demoulding transferred to airtight plastic bags for further curing at 38°C for 27 days before
being tested. Reference cubes of OPC without replacement, were also cast. Mini-RILEM
2 × 2 × 15 cm 3 mortar prisms (F.L. Smidth Denmark A/S, 1969) were also tested to study
168 C. He et al. / Applied Clay Science 9 (1994) 165-187

the development of compressive strength with age. The mix ratio.


OPC:(meta)kaolin:sand:water, was 0.7:0.3:3.0:0.6. The prisms were cured for the first day
as for the ASTM test. After demoulding they were transferred to water for further curing al
40°C tbr 2, 7, 28 and 91 days, when tests were carried out. Each prism was sawn into 3
pieces and the mean value of three tests was normalized to a reference water and air content
using Feret's formula (Neville, 1981 ).
Hardened cement mortar specimens and reacted kaolin pellets were vacuum dried. Scan-
ning electron microscopy (SEM) was performed with a Philips SEM 515 microscope on
gold-coated specimens. Electron microprobe analysis (EPMA) was carried out with a JEOL
Superprobe with TRACOR energy dispersive system.

3. Results and discussion

3.1. Mineralogical characterization of kaolin

The chemical and mineralogical composition of the kaolin sample is shown in Table 2.
The sample consists overwhelmingly of SiO2 and A1203 (81.4%) with minor amounts of
K and traces of other components. The K20 content of the sample corresponds to the
admixture of 12.0 wt% muscovite and 2.3 wt% microcline.
The XRD patterns, both from the bulk sample and its clay fraction ( < 2/zm) (Fig. 1
and Fig. 2 respectively) indicate predominance of well-crystallized kaolinite with a little
associated mica, quartz and minor microcline. The crystallinity index of kaolinite, measured
and calculated according to the method of Hinckley (1963), is 1.10 indicating relatively
well-crystallized material.
Fig. 3 shows the DTA and the H20 release curve. No release of CO2 was detected. The
endothermic reaction at 565°C, accompanied by a large H20 release, indicates the dehy-
droxylation reaction of kaolinite. In the DTA and TG curves (Fig. 4) the small endothermic
peak below 100°C represents release of 1.62 wt% of absorbed water. The extraordinarily
curved convex DTA base line that follows this release is probably due to heat-conduction
problems in the sample after dehydration. Dehydroxylation of kaolinite gives a pronounced
endothermic peak at 570°C together with a weight loss of 11.3%. The difference in tern-

Table 2
Chemical and mineralogical composition of studied kaolin (wt%)

SiO2 A1203 Fe203 FeO K20 Na20 TiO2 P205 MgO CaO Volatile

46.1 35.3 1.01 0.05 1.62 0.06 0.05 0.12 0.27 0.06 12.11
(46.6 39.5 . . . . . . . . . 14.0) a

kaolinite mica quartz microcline crystallinity index

82 12 3.8 2.3 1.10

"Chemical composition of ideal kaolinite.


C. He et al. /Applied Clay Science 9 (1994) 165-187 169

F- Microcline
M - Mi¢o
13 - ~ J o r t z
K and unmorked peolot - Koollnlte
100

80 1
M iI (M)

2O

"" ~ . . . . lb . . . . l~ . . . . 2b. . . . ~ . . . . ~b. . . . • . . . . +b. . . . + . . . . • . . . . + . . . . ell"


2O
Fig. 1. XRD pattern of the bulk kaolin sample.

peratures between the two instruments is negligible. A sharp exothermic peak due to
recrystallization started at 960°C and culminated at 980°C. The total loss on ignition, 12.9
wt%, agrees well with the bulk chemical analysis.
SEM (Fig. 5) shows that kaolinite crystals occur as packets o f thin flakes about 1-5/+m
across. Typical kaolinite " b o o k s " were sometimes observed. Individual flakes have pseudo-
hexagonal or irregular, or multiply lobed margins. Very few perfect hexagonal kaolinite
crystals were found.

3.2. Calcination effects

Destruction of kaolin by dehydroxylation and formation of ' 'metakaolin" can be achieved


by heating between 550°C and 700°C (Mackenzie, 1970; Babushkin et al., 1985). The
CD

¢,i • O

<5
;oo. ,~
qP

80

60
/
+t
'5 .... Ib" " "I~ .... 2'o'
2e
Fig. 2. XRD pattern of the clay fraction ( < 2 p,m) of the studied kaolin sample.
170 C He et a L / Applied Clay Science 9 (1994) 165-1,~7

565°C

H20

565°C
7OoC /~

....... ". . . . . . . . . . . . . . . . . . . . ~-~-~D---"


JTA "-~ "- ----~--

Fig. 3. DTA and volatilereleasecurvesfor the studiedkaolinsample.

optimum temperatures for calcined kaolin as pozzolan were reported to range from 6 4 8 ° C
to 982°C (Mielenz et al., 1950), from 538°C to 932°C (Price, 1975), from 680°C to 980°C
(Gao and Zhang, 1989), or 850°C (Ayub et al., 1988). In the present study, four calcination
temperatures were selected according to the thermal behaviour of the samples revealed by
DTA and TG. They are ( 1) 550°C, the culmination point of the endothermic dehydroxy-
lation reaction; (2) 650°C, at which this endothermic reaction is completed; (3) 950°C,
the starting temperature of recrystallization; and (4) 800°C, an intermediate temperature
between (2) and (3). The heating time needed to achieve reaction completion is a function
of several factors, such as the heating rates and packing mode, as well as the intrinsic
properties of kaolin, such as its crystallinity and grain size distribution. XRD patterns of
the calcined kaolin samples in the present investigation showed that almost all kaolinite
peaks disappeared after calcination at 550°C for 40 minutes. The DTA curve also showed
that the main endothermic reaction is completed at about 650°C with a temperature rise of
10°C per minute. Accordingly, an annealing time of 100 minutes was used throughout the
present study.
X R D p a t t e r n s for the samples annealed at 550°C, 650°C, 800°C and 950°C for 100 minutes
are shown in Fig. 6. The most striking is their enormously enhanced general background

TG .,'"
-0.00

o.O2

o3
).04 0

~/i/, c-
Cr~
\j' •o . o 6 'g

[ i t i * i i

1200 1000 860 600 460 260 o


temperature (*C)
Fig. 4. DTA and TG curvesfor the studiedkaolin.
C. He et al. /Applied Clay Science 9 (1994) 165-187 171

Fig. 5. SEMphotographsof kaolinsample(goldcoated).

caused by the amorphous material in the samples. The intensity of the diffuse band in the
2 0 region from 20 to 30°, the highest region of general background, increased by 78 percent
after the raw sample was calcined at 550°C. It further developed with temperature, up to
95% at 800°C but dropped abruptly to 43% at 950°C (Fig. 7a) due to recrystallization.
After calcination at 550°C, kaolinite peaks at 7.25/~, 3.58/~ and 2.78/~, are still detectable
although very weak. The residue of kaolinite, using the well preserved mica peaks as the
internal intensity standard, is about 1.1% of the total original amount. In the XRD pattern
of the samples calcined at 650°C and 800°C all kaolinite peaks completely disappeared. In
the pattern of the sample calcined at 950°C, a new set of broad peaks of mullite, at 5.41 ,~,
3.42/~, 2.70/~, 2.29/k and 2.21/~ appear (Fig. 6). Thus the X-ray diffraction analysis
indicates that only remnants of kaolin persist after annealing at 550°C for 100 minutes; pure
metakaolin is obtained from kaolin during the calcination process at 650 and 800°C; at
172 C. He et al. /Applied Clay Science 9 (1994) 165-187

M
0
M - mica
I Q - quartz
K - kaolinite
II F - microcline
II Mu - 'mullite

K "|IF M M

tff,~c

950°C 7
Mu
10 15 20 25 30 35 40
2e
Fig. 6. XRD patterns of calcined kaolin. Room temperature measurementsafter annealingfor 100 minutesat
indicatedtemperatures.

950°C recrystallization and formation of mullite starts. Mica, quartz and microcline are still
observable after 100 minutes at 950°C.
In the S E M p h o t o g r a p h s of calcined samples (Fig. 8), there is no significant difference
between the 550°C sample and untreated kaolin (Fig. 5 ). Besides the multiply lobed flakes,
kaolinite crystals with hexagonal outlines were found occasionally preserved in the sample
calcined at 550°C, in agreement with its XRD data. The kaolinite flakes in the samples
calcined at 650°(3 or 800°C are more deformed and locally condensed into bundles. Irregular,

Fig. 7. Changes in physical and chemical properties of kaolin on calcination.(a) Enhancementof the diffuse
background of XRD patterns of metakaolin,in % of its intensityincreasein the band 20= 20-30° over that of
unheated kaolin; (b) density; (c) specific surface area; (O) Blaine, ([3) BET. (d) AlkalisolubleSi and A1 in
%0of the originalsampleweight; ([]) A1, (O) Si, (A) Si + A1. (e) Acid solubleSi and AI; same symbolsas for
(d).
C. He et al. /Applied Clay Science 9 (1994) 165-187 173

200- 2850-
~e

/ x ¢
c 180
"(3( ~ ,.-,,2800- iI
.(3
I
E I
160- x
iI

"-"2750- I
GL /
",, ,,~
X ", #

. ~

i
2700-
1201
\ I

2650
i00 260 ~60 ' 660 ' 860 ' ~obo ' 260 ' 460 ' 660 ' 860 1obo
T r e a t m e n t Temp. C T r e a t m e n t Temp. C
a b

1600- -17 80 =
,A
-t6 / \\
/
1400, ~ 60, / \
3" / \
/ \ -15 o
/ " ~ ~l vE
/ /
/
,,~E1200- .~_ 40. l ~e - -Q
/ / XX 0
/

o
/ Xi // ,s/~ , \
m
/ / ,./.." \
/
1000 .~ 20" // /~,/ \
/ // // \\
~ "12~
<

800
26o ' 460 ' 66o " 8 N '~ooo 2(~0 ' 460 ' 660 ' 8~0 ' 10bO
T r e a t m e n t Temp. C Treatment Temp. C
0 d

2.5 A-A~

/ t
2.0
/ / X
0 / / %A
/ / x x
,2-01.5 / / \ x
/ / X
/ / / XXX
/ \A
"~ 1.0 - //

"~ 0.5,
< . O _ _O.. - - O -

0.0
260 ' 46o ' 66o ' 86o 'lobo
T r e a t m e n t Temp. C
e
174 C He et at./Applied Clay Science 9 (1994) 165-187

Fig. 8, SEM photographs of calcined kaolin from the following temperatures: (a) 550c'C, ( b ) 650°C. ( c ) 800c(
and (d) 950°C

multiple splitting of thicker flake packets is observed at 550, 650 and 800°C. After calci-
nation at 950°C the sample appeared more massive and flakes had sintered rims. The
dehydroxylation of kaolinite seemed not to affect the external shapes of flakes very much;
this corroborates the observation of Brindley and Lemaitre (1987) that a degree of structural
order persists.
Particle size distribution of a pozzolan influences proIbundly its pozzolanic activity.
Kaolin calcined to different temperatures and subsequently subjected to the same grinding
process turned out to have different density and fineness (Table 3, Fig. 7b, c, Fig. 9). The
10% weight toss due to dehydroxylation as well as the multiple splitting of kaolinite flakes,
led to a 3% decrease in overall density but a 54% increase in specific surface area (Blaine)
Table 3
Density and specific surface area of calcined kaolin

Analyses Method Unit Untreated Samples calcined at

550°C 650°C 800°C 950°C

Density Gas. Pyc. kg/m 3 2740 2660 2700 2730 2820


Spec. surface Blaine mZ/kg 983 1515 1422 1280 818
Spec. surface BET mZ/g 15.3 16.6 16.4 15.9 11.2
C. He et al. /Applied Clay Science 9 (1994) 165-187 175

45-
.---ll--

40- Untreated
O

>..35- 550°C

~- 3 0 - 650oc
(1)

25 800°C
c-
O
-,-- 20 950°C
..(3

j
o_
~- 15
-4--
,~
E3 10

0.1 1 10 10(] 1000


Particle Size (pm)
Fig. 9. Particlesizedistributionof calcinedkaolin.

of kaolin following 550°C dehydroxylation. Further heating leads to slow compaction of


layered crystals, indicated by increasing density and decreasing specific surface. The density
surpasses that of untreated kaolin by 3% and the specific surface area drops 17% below the
original values as sintering and growth of new crystal phases occur at 950°C (Fig. 7b, c).
Particle size distribution for the ultrasonically dispersed sample displays a broad peak in
the 3-10/zm range for the untreated kaolin and for those treated at 550-650°C. With
development of sintering and recrystallization, another particle concentration peak develops
in a coarser region, around 25/zm for 800°C and around 70/zm for 950°C (Fig. 9).
Measurement of alkali-soluble silica has been suggested as a simple, rapid method for
evaluation of pozzolanic activity (Surana and Joshi, 1990). However, we observed that the
soluble alumina values agree better with the assessment of pozzolanic activity obtained by
other methods, to be discussed below. After kaolin had been calcined to 550°C, the amounts
of alkali-soluble silica and alumina respectively increased to 545% and 630% of the original
values. They reached a peak of 595% and 650%, respectively, at 800°C and decreased again
at 950°C (Table 4 and Fig. 7d). The 3.8-fold decrease in soluble alumina from 800°C to
950°C is abrupt but no corresponding decrease is observed for soluble silica. This suggests
that the striking reduction in the amount of amorphous matter between 800 and 950°C,
revealed by XRD diffraction (Fig. 7a), occurs with consumption of AI as mullite crystal-
lizes. We propose that for the pozzolanic materials rich in alumina, such as most clays, the
alkaline solubility of alumina is a valid measure of pozzolanic activity.
Acid solubility o f silica and alumina from pozzolanic metakaolin was also studied in the
present investigation. The results agree generally with other indications of pozzolanic
activity, and again, soluble alumina is a more sensitive indicator of the liability to acid attack
176 ('. He et ~i./ Applied Clay Science 9 (1994) 165-187

Table 4
Alkali and acid solubility of untreated and calcined kaolin

Calcination temp. Reagent NaOH (ppm) Reagent HCI (ppm)


(°C)
Si AI Si+ AI Si AI Si+ AI

20 6650 5420 12,070 250 824 1074


550 36,350 34,190 70,540 390 2056 2446
650 35,050 32,060 67,110 370 2083 2453
800 39,550 36,035 75,585 440 1838 2278
950 40,550 9540 50,090 240 893 1133

than silica (Fig. 7e). Besides, comparison of Fig. 7e with Figs. 7a, b and c indicates that
the acid solubility test is a very sensitive and simple indicator of disordering and/or recrys-
tallization of kaolin on calcination; both of these determine the pozzolanic activity.

3.3. Reaction of calcined kaolin pellets with SCPS

To study this reaction, calcined kaolin pellets were immersed in SCPS. Observations
were undertaken on the samples that had been immersed in the solution for 7, 28 and 90
days, respectively, at 50°C.
After 7 days, reaction products were already visible by SEM on the pellets from all
calcination temperatures. White "dusting" appeared on the surface of untreated kaolin
(Fig. 10a). Flower-like aggregates with morphology typical of gehlenite hydrate were
found growing on the surface of 550°C sample (Fig. 10b) and gel-like phases on the 650°C,
800°C and 950°C samples (Fig. 10c, d, e). Only sporadic mineral fibers were observed in
untreated kaolin after 28 days (Fig. 1 la); reactivity of the calcined samples was consistently
higher (Fig. 1 lb-e). A coating of the surface by reaction products (Fig. 12) hinders further
reaction. The effect of precipitation of CH on the pellet surface and crystallization of calcite
due to atmospheric carbonation is shown in Fig. 10b.
To investigate the reaction products, surface and sub-surface layers of reacted pellets
were scraped away for XRD analyses. Except for the peak of quartz at 3.35 .A the loose
surface layers are mainly amorphous material. Fig. 13 shows some XRD patterns of sub-
surface layer, with a pattern of unreacted 550°C metakaolin added for comparison. It is
obvious that no appreciable amounts of XRD-detectable products were produced during the
immersion. A weak peak of calcium carbonate was detected for each sample, probably due
to atmospheric carbonation. This peak is strongest for the sample calcined at 800°C and
after 28 days immersion. The same sample also displays two medium strong peaks at 7.61
and 3.80 ,~.,attributable to the basal spacings 0001 and 0002 of calcium monocarboaluminate
hydrate (C4A(~H12) !
Electron microprobe was employed to investigate pellets immersed for 7 and 28 days.
Even though attention was directed towards newly formed products, the compositional
points predominantly cluster in the kaolinite, mica and microcline areas (Fig. 14) with

qn cement chemistry, the following abbreviations are generally used: C = CaO, A = A1203, S = SiO2, H = H20,
= SO3, C = CO2, etc.
C. He et al./ Applied Clay Science 9 (1994) 165-187 177

Fig, 10. Fracture surfaces of kaolin pellets reacted with SCPS for 7 days. (a) Unheated sample; (b) sample
calcined at 550°C; (c) at 650°C; (d) 800°C and (e) 950°C.

C2ASH 8, C4AHx and their mixtures in the samples from 550°C with 7 days immersion and
800°C, 28 days immersion. Slightly aluminous C - S - H with a C:S ratio of about 1.2 was
occasionally detected.
Evidently, reaction between CH and calcined kaolin pellets in SCPS is slow. Compression
of powdered kaolin on pelletizing appears to affect adversely liquid penetration; a precipitate
of CH and of reaction products covers the surface and hinders further reaction.
178 C He eta/. / Applied Clay Science 9 (1994) 165-187

Fig. I I. SEM photographsof fracture surfacesof calcined kaolin pellets reacted with the SCPS for 28 days. (a)
Unheated; (b) calcined at 550°C; (c) at 650°C~ (d) 800°C and (e) 950°C.

Physical and chemical properties o f (meta)kaolin-cement mortars


The results of compressive strength tests (Tables 5 and 6, Fig. 15a, b) showed that the
mortars with 20 wt% replacement of OPC by calcined kaolin have much higher strengths
than the mortar containing uncalcined kaolin and higher than the mortar with pure OPC (all
comparisons refer to the situation after 28 days curing at elevated temperature; for details
see Section 2). From compressive strength alone, the optimum heating temperature is 800°C,
with 31% increase of compressive strength over pure OPC. The results for 550°C and 650°C
are somewhat lower than that for 950°C but still higher by 17-20% than for OPC (Table 5
and Fig. 15a).
C. He et al. /Applied Clay Science 9 (1994) 165-187 179

Fig. 12. Amorphousreactioncoatingon the surfaceof kaolinsamplesimmersedin SCPS. (a) Calcinedat 650°C,
immersedfor 7 days, (b) at 800°C,immersedfor 90 days.

The tests of Mini-RILEM prisms ( see Section 2) also indicated, for all curing ~imes from
2 days to 91 days, the enhancement of compressive strength of kaolin-cement mortars by
calcination of their kaolin component (Table 6 and Fig. 15b). The strength of the mortar
with kaolin pre-calcined at 950°C was low after 2 days but increased quickly with curing
time and reached the same level as other activated specimens by 91 days. Evidently,
calcination of kaolin to the temperature range between 550 to 950 increases greatly the
compressive strength of their mortars; after curing for minimum 7 days, the results for these
temperatures are essentially comparable.
The hydration reaction of cement is accompanied by a reduction in the total volume of
the system (Geiker and Knudsen, 1982; Taylor, 1990). According to Knudsen (1985),
Geiker's method for the measurement of this effect, the so-called chemical shrinkage can
be used to evaluate the rate and degree of pozzolanic reaction. Both the mixtures kaolin-
180 C, He et dl. / Applied ('lav Science 9 (1994) 165-1~¢7

Q
M

M
Q
K
-
-
-
mica
quartz
kaolinite
f!
F - microcline
C - calcite ii[
A - calcium aluminate M KFI ' M M
carbonate hydrate Q A•I ~ • M K M

c ., l I

ci I/LI f .....

c, "

1 , , I ~ ' ' ' I ,- , , ~ T - - ~ - , I ~ " ~ ~ ~ ' , --y~---r--~ , ~~ " ~


5 10 15 20 25 30 35 40

2(3

Fig. 13. XRD patterns of unreacted kaolin and of the subsurface layer of pellets reacted with SCPS. Symbols used
indicate calcination temperature in °C/length of immersion in days.

CH and kaolin-OPC were tested in this way in the present investigation (Fig. 15c, d).
After a short initial period, the pastes of (meta)kaolin from 550°C, 650°C or 800°C
mixed with CH powder and SCPS (for details see Section 2) reacted faster than the pastes
with kaolin uncalcined or calcined at 950°C. The pastes with 550 and 650°C metakaolin
exhibit larger chemical shrinkage than reference fly ash. For the mixtures of (meta)kaolins
with OPC, the specimens with kaolin uncalcined or calcined at 950°C persistently showed
poorer reactivity (Fig. 15d); the remaining mixtures gave results comparable with fly ash.
The shrinkage of samples hydrated for 45 days versus calcination temperature is depicted
in Fig. 15e; this figure demonstrates clearly the increase in the pozzolanic activity of kaolin
after calcination between 550 and 800°C.
Since most of the cement hydration products are poorly crystallized and for many phases
X-ray diffraction peaks overlap, the difficulties in determining the mineralphases in reacted
mortar are well known (Midgley, 1962; Taylor, 1990). Fig. 16 shows XRD patterns of
hardened cement mortars with and without admixtures of (meta) kaolin (preparation details,
Section 2). The principal lines belong to CH (especially for poor pozzolan) and to quartz
introduced as sand. The main reaction product, tobermorite (C586H9) or C-S-H (I), with
a diffuse band in the region 3.05 to about 2.78 .~ and a somewhat sharper peak at 1.82 A is
always present. Gehlenite hydrate (C2ASH8), calcium aluminate monosulphate
C. He et al. /Applied Clay Science 9 (1994) 165-187 181

• I-ITIT1 K550~28
Ks oz28

AI203 CoO

Fig. 14. Electronmicroprobeanalyses (in mol%) of calcinedkaolinpelletsimmersedin the SCP S for 7 daysand
28 days. Symbolsindicate calcinationtemperaturein °C/length of immersionin days.

(C4ASHx), as well as calcium aluminate hydrate, C4AHI3, were also detected. Remnants
of the slowest reacting clinker mineral,/3-C2S, are found in all specimens. The kaolinite
peaks, 7.12 and 3.58 fi,, are present in the XRD chart of the mortar containing untreated
kaolin, but not in the specimens with the calcined kaolin.
SEM was employed to investigate the new reaction products and the micro-structure o f
hardened mortars. An amorphous gel phase was found in every specimen (Fig. 17). Lath-
shaped tobermorite phase was observed in the specimen containing kaolin treated at 800°C
(Fig. 17d). A fibrous C - S - H phase was found to develop in the specimen mixed with
kaolin calcined at 950°C (Fig. 17e); this might partly explain its rather high compressive
strength. The kaolinite residues with flaky shapes that might be deleterious to the strength
were common in the specimen containing untreated kaolin (Fig. 17a).
Fresh m o r t a r f l o w is an important practical technical parameter; its adjustments affect
subsequent mortar strengths. Calcination of kaolin at increasing temperatures steadily
increased the flow of resulting mortars (measured according to ASTM C230 and C109)

Table 5
Compressivestrength of cement-kaolinmortars (ASTM)

Mortars OPC (Ref.) Kaolin additive calcinedat followingtemp.:

20oc 550°C 650°C 800°C 950°C

MPa 34.99 26.31 41.95 41.11 45.80 43.61


% of OPC 100 75 120 117 131 125

Each value is an averageof 3 measurements;the mean (maximum)coefficientof variationis 3.2% (7.2%). 20


wt% replacementof OPC by untreated/calcinedkaolin.
182 (', He et af / Applied Clay Science 9 f 1994) 165-1~'7

140 140-
"S .0
~'130-
U3
< 1 120~ -z,
~120- / ~ / //[3- ...... [~.
%

~3
I0- / ~1oo4
~2 /

~I00- z/ gO~ // / / / / ~ \
\
CO g o I //
I / / ///
70 j ,, /. ". 'o
80]D,// I I ii

5
~£ 7040 ' 260 460
' 660 ' 860 1000 ~. so ' 260 ' 46o 86o ' 86o '~obo
Treatment T e m p . C Treatment Temp. C
a
b

7.0 7.0
( ~ cement
6.0 6.0

05.0 0~5.0 ~ ' , : ; I gso

-E4.0 --'~"E4..0

~3.0 ~3.0 -
c E
'c 2.0 "~2.0
.C
(.n 03
1.0 1.0~

0.0
1 10 100 1000
00J 1 I0 100 1000
t (hour) t (hour)
c d

7.0 120"

o
/ P
u-~ 6.0 / \
/ \ I
v / \
/
, ) 100- iI

0 / \ iI
0 // \
.~5.0 / \
/ \ 0 //
// \ h
/ \ 80
/
s~
\
~4.0
E

3.0
• 260 ' 4 ~ ' ~ ' 860 10'00
60
260 ' 460 ' 660 ' 860 ' 1obo
Treatment Temp. C Treatment Temp. C
e f
C. He et al. /Applied Clay Science 9 (1994) 165-187 183

Table 6
Compressive strength of cement-kaolin mortars (Mini-RILEM)

Temp. MPa %

(°C) 2d 7d 28d 91d 2d 7d 28d 91d

20 18.34 27.99 37.50 37.89 54 52 54 58


550 30.74 54.66 78.30 80.51 91 102 113 123
650 32.53 62.10 84.58 83.27 96 116 122 127
800 32.48 61.92 81.45 80.51 96 115 118 123
950 23.88 57.31 84.30 87.24 71 107 122 133
OPC(Re~) 33.73 53.74 69.24 65.56 100 100 100 100

Each value is an average of 3 measurements;the mean (maximum) coefficientof variation is 1.7% (4.2%). 30
wt% replacement of OPC by untreated/calcined kaolin

and most evidently at 950°C (Fig. 15f). Hence, the amount of water necessary to obtain
the same flow in the mortars decreased in the test prisms, this contributing to their higher
compressive strength (Lea, 1970).

4. Conclusions

The three major chemical components, SiO2 + A1203 + Fe203, constitute 82.41 wt% of
the investigated kaolin. This conforms with the requirements ( > 70%) stated in A S T M
C618-89a for a natural pozzolan. Mineralogically, the sample is composed of 82% kaolinite,
12% mica, 3.8% quartz and 2.3% feldspar by weight.
Our study confirms that calcination is an effective approach to enhance the pozzolanic
activity of kaolin. Most of kaolinite is destroyed and altered to an X-ray amorphous phase
after heating at 550°C for 100 minutes. The amount of XRD amorphous metakaolin increases
modestly during further calcination to 650°C and 800°C but starts to fall by 950°C when
mullite appears. This amount is effectively measured by the background intensity of the
XRD patterns.
Results of chemical shrinkage for the mixtures of ( m e t a ) k a o l i n - O P C or ( m e t a ) k a o l i n -
CH show that the kaolins treated at 550°C and 650°C have the fastest reaction rate, slightly
higher than the reference fly ash. They are followed by kaolin calcined at 800°C. The faster
reaction rates are in agreement with the higher specific surface of 550-650°C metakaolin.

Fig. 15. Changes in technological characteristics of pozzolaniccement with calcinationtemperature of the kaolin
component. (a) Relative compressive strengths of cement- ( meta) kaolin mortarsin % of the strength of reference
OPC. 20 wt% of OPC replaced by (meta)kaolin; 28 days at 38°C. (b) Relative compressivestrengthsof cement-
(meta)kaolin mini-RILEM mortars in % of the strength of reference OPC. 30 wt% replacement. (O) 2 days;
([]) 7 days; (A) 28 days; (<>) 91 days; at 40°C. (c) Chemical shrinkage of CH-(recta)kaolin 1:1 mixtures.Fly
ash-CH mixture was used as reference. (d) Chemical shrinkage of OPt:-(meta)kaolin 1:1 mixtures. Pure OPC
and OPC-fly ash mixture were used as reference. (e) Chemical shrinkage of OPC-(meta)kaolin l:l mixtures;
measured on 45th day. (f) Flow of OPC-(meta)kaolin mortars with 20 wt% replacement. Water/
(cement+ (meta)kaolin) ratio is 0.6.
184 (:, He et a~ /Applied Clay Science 9 (1994) 165 187

Q~

K - Kaolinite Q - Quartz
c - C-S-H B - C2S
Ca - Ca(OH)2 A - C4AH13 Ca
C
S - C4A'~Hx T - C2ASH 8
A
Ca

Q 1 Ca KK

K KI
S
raw kaolin

550 C

650 C

800 C

950 C

''"5 .... ~b . . . . ~'5 . . . . 2'o . . . . 2'5 . . . . 3'o . . . . 3'5 . . . . . 4'o . . . . ;s .... 5~ . . . . s'5 . . . . 6'o'"

Fig. 16. XRD patterns of hardened cement-(meta)kaolin mortars cured at 38°C in the air for 28 days. 20 wt% of
OPC replaced by kaolin untreated or calcined at 550, 650, 800 and 950°C, respectively
C. He et al. /Applied Clay Science 9 (1994) 165-187 185

~ 4

Fig. 17. SEM photographs of fracture surfaces of hardened cement-kaolin mortars cured at 38°C in air for 28
days. 20 wt% of OPC was replaced by (a) unheated kaolin or by kaolin calcined at: (b) 5500C, (c) 6500C, (d)
800°C and (e) 950°C. (f) OPC without kaolin.

Measurements of alkali and acid soluble Si and A1 for (meta)kaolin correlate well with
the results of technological tests. If the tested material is rich in alumina, alkali soluble AI
can be more informative about its pozzolanic activity than soluble silica. Acid solubility is
a readily available test for structural decomposition and new phase formation during cal-
cination.
The main reaction products of the hardened (meta)kaolin-cement mortars are same: C -
S-H, C2ASH8, CaASHx and C4AH13,but they occur in various quantities. They are less
186 c. He et al / Applied (?lay Science 9 (1994) 165-o187

pronounced in the mortars containing kaolin untreated and treated at 950°C, in full accord
with their low content of amorphous phase and low chemical shrinkage. SEM studies of
the paste show that the mixtures from 800°C and 950°C display the best intergranular
network of hydrated phases.
The technological tests on both mortar cubes and mini-RILEM prisms indicate that
calcination of kaolin in the temperature range between dehydroxylation and recrystallization
increases the compressive strengths substantially. All tested mortars with kaolin calcined
in this range have relative compressive strengths much higher than 75% of that of the OPC,
the threshold value stated in the ASTM C618-89a as the minimum requirement for a
pozzolan.
Our study points out that the improvements in most of the structural and technical
properties of the mortars are modest for calcination temperatures from 550 to 800°C, the
primary reason being that the increase in structural disorder in this temperature range is
minor compared to that caused by dehydroxylation at 550°C. The positive influence of
structural disorder is partly counterbalanced by the decrease in specific surface area and
increasing grain size which, however, result in better flow values. Mixtures from 950°C can
attain compressive strength comparable with that from 550 and 800°C but they have a very
low reaction rate.
Adding the economic point of view for the mass production of metakaolin pozzolans to
the technological considerations, calcination at 550°C in sufficiently thin layers yields a
good pozzolan at modest time expenditure.

Acknowledgements

The authors wish to thank Assoc. Prof. J. Rensbo and Mr. J. Flyng for the keen interest
and assistance at the electron microprobe analyses and Assoc. Prof. E. Leonardsen and Mrs.
T. Poulsen for the X-ray diffraction analyses. Mr. H.B. Jensen in F.L. Smidth and Co. A/
S kindly provided assistance in the technical tests. We thank Dr. T. Knudsen at Danish
Technical University for his assistance in the chemical shrinkage tests and to him and an
unknown referee for constructive discussions. Material and apparatus needed for the project
were financed by the University of Copenhagen, F.L. Smidth and CO. A/S and Danish
Natural Research Council. Kind assistance of Assoc. Prof. J. Rensbo was instrumental in
setting up this project.

References

ASTM C618-89a, C311-90, C109-88 and C230, AmericanSocietyfor Testingand Materials.


Ayub, M., Yusuf, M., Beg, A. and Faruqi, F.A., 1988. Pozzolanicpropertiesof burnt clays. Pak. J. Sci. Ind. Res.,
31(1): 1-5.
Babushkin, V.I.,Matveyev,G.W. and Mchedlov-Petrossyan,O.P., 1985.Thermodynamicsof Silicates. Springer,
Berlin, pp. 165-167.
Brindley, G.W. and Lemaitre,J., 1987. Thermal,oxidationand reduction reactionsof clay minerals. In: A.C.D.
Newman (Editor), Chemistryof Claysand Clay Mineralogy.MineralogicalSociety,LongmanScientificand
Technical, pp. 319-370.
C. He et al. /Applied Clay Science 9 (1994) 165-187 187

Cook, D.J., 1985. Calcined clay, shale and other soils. In: R.N. Swaing (Editor), Cement Replacement Materials.
Concrete Teehnoi. Design, 3: 40--72.
De Silva, P.S. and Glasser, F.P., 1992. Pozzolanic activation of metakaolin. Adv. Cement Res., 16:167-178
De Silva, P.S. and Glasser, F.P., 1993. The phase relation in the system CaO-A1203-SiO2-H20 relevant to
metakaolin --calcium hydroxide hydration. Cement Concrete Res., 23:627--639
European Standard EN196, 1987.
F.L. Smidth Denmark A/S, 1969. Report on the Mini-RILEM-method. Internal Report APr/EP.
Forrester, J.A., 1975. Burnt clay pozzolans. Proc. Meeting on Small-scale Manufacture of Cement Materials.
Intermed Technol. Publ., London, pp. 53-59.
Gao, Q. and Zhang, Z., 1989. Study on the structure change in the calcination process of kaolinite and its
pozzolanicactivity. J. Chin. Ceram. Soc., 17(6): 541-548 (in Chinese, with English abstract).
Geiker, M. and Knudsen, T., 1982. Chemical shrinkage of portland cement pastes. Cement Concrete Res., 12:
603--610.
Hinckley, D.N., 1963. Variability in "crystaUinity values" among the kaolin deposits of the coastal plain of
Georgia and south Carolina. Clays Clay Miner., 13: 299-235.
Knudsen, T., 1985. On the possibility of following the hydration of fly ash microsilica and fine aggregates by
means of chemical shrinkage. Cement Concrete Res., 15: 720-722.
Lea, F.M., 1970. Chemistry of Cement and Concrete. Arnold, London, pp. 171-245.
Mackenzie, R.C., 1970. Differential Thermal Analysis, 1. Academic Press, London, pp. 524-529.
Midgley, H.G., 1962. The mineralogical examination of set Portland cement. 4th ISCC, Washington, 1: 479-490.
Mielenz, R.C., Witte, L.P. and Glanz, O.J., 1950. Effect of calcination on natural pozzolans. Symp. on Pozzolanic
Materials in Mortars and Concrete. ASTM, pp. 43-92.
Murat, M., 1983. Hydration reaction and hardening of calcinedclays and related minerals, I. Preliminary investi-
gation on metnkaolinite. Cement Concrete Res., 13:259-266
Neville, A.M., 1981. Properties of Concrete. Pitman, London, 268 pp.
Price, W.H., 1975. Pozzolans - - a review. ACI J., 5: 225-232.
Serry, M.A., Taha, A.S., E1-Hemaly, S.A.S. and E1-Didamony, H., 1984. Metakaolin-lime hydration products.
Thermochim. Acta, 79:103-110
Surana, M.S. and Joshi, S.N., 1990. Estimating activity ofpozzolanic materials by spectrophotometric method.
Adv. Cement Res., 10: 81-83.
Taylor, H.F.W., 1990. Chemistry of Cement. Academic Press, New York.

Das könnte Ihnen auch gefallen