You are on page 1of 90

_, SHRP-C-628

Concrete Microstructure
Porosity and Permeability

D.M. Roy
P.W. Brown
D. Ski
B.E. Scheetz
W. May

Materials Research Laboratory


The Pennsylvania State University
University Park, Pennsylvania

Strategic Highway Research Program


National Research Council
Washington, DC 1993
SHRP-C-628
Contract C-201

Program Manager: Don M. Harriott


Project Manager: lnam Jawed
Production Editor: Marsha Barrett
Program Area Secretary: Ann Saccomano

April 1993

key words:
cement paste
concrete
permeability
pore structure
pore size distribution
porosimetry
porosity
surface area

Strategic Highway Research Program


National Academy of Sciences
2101 Constitution Avenue N.W.
Washington, DC 20418

(202) 334-3774

The publication of this report does not necessarily indicate approval or endorsement of the fmdings, opinions,
conclusions, or recommendations either inferred or specifically expressed herein by the National Academy of
Sciences, the United States Government, or the American Association of State Highway and Transportation
Officials or its member states.

© 1993 National Academy of Sciences

350/NAP/493
Acknowledgments

The research described herein was supported by the Strategic Highway Research
Program (SHRP). SHRP is a unit of the National Research Council that was authorized
by section 128 of the Surface Transportation and Uniform Relocation Assistance Act
of 1987.

iii
Contents

Acknowledgments .................................................. iii

Abstract .......................................................... vii

Executive Summary .................................................. 3

A Model for the Distribution of Pore Sizes in Cement Paste .................... 7

Lognormal Simulation of Pore Evolution During Cement and Mortar Hardening ... 19

Concrete Microstructure and Its Relationship to Pore Structure, Permeability,


and General Durability ............................................ 25

Porosity/Permeability Relationship ...................................... 43

Relationship Between Permeability, Porosity, Diffusion and Microstructure


of Cement Pastes, Mortar, and Concrete at Different Temperatures ........... 76

/4

V
Abstract

A model has been developed that lays the foundation for relating porosity to
permeability. This is based on knowledge gained from previous work as well as
experimental and theoretical input from the present program. A linear combination of
lognormal distributions may be used to define the pore structure. This report contains
five papers relevant to this topic.
EXECUTIVE SUMMARY

The objective of the current study was to develop an accurate mathematical descriptor for
pore size distribution to relate to permeability and use for permeability prediction. Various
properties of cement-based materials are affected not only by total porosity but also by the size
distribution of the porosity present. In order to model the relationship between pore size
distribution and properties, a suitable mathematical descriptor for pore size distribution must
first be found. Mercury intrusion porosimetry (MIP) is a commonly used method of determining
pore size distributions for the range of pore sizes which significantly affect properties such as
permeability. A suitable mathematical descriptor for pore size distributions determined by MIP
should be one that not only fits the experimental data but, more importantly, provides the basis
for the physical interpretation of pore structure.
Diamond and Dolch (Ref. 1, Paper 1, this supplemental report) showed that the pore size
distributions in cement pastes determined by MIP could be described as lognormal distributions.
The present work in which higher intrusion pressures have been applied, suggests that a single
lognormal distribution is inadequate to describe the smaller pores. A mixture of two lognormal
distributions, or a compound Iognormal distribution was suggested to model the distribution of
pore size determined by MIP in both cement pastes and mortars. However, closer examination of
the data used in our model shows that very coarse pores, pores - > 1 ]an, were rarely included. Pores
in this size range affect properties, such as strength and permeability, and should not be ignored.
In the present study, pores as large as 7 _n are included. It appears that a mixture of three
Iognormal distributions can more closely fit the distribution of pore sizes over the range of sizes
extending 4 orders of magnitude from about I nm to about 10 _rn.
A variate X (0 < x < _) is Iognormally distributed if Y = IogX is normally distributed with mean
and standard deviation ¢_.The probability density function is,
P(X) = (2r_s2)'1/2X" lexp[-0.5{flog(X) - _)/¢_}2] [3]
The range of X is 0 < X < _. _ is defined as the location parameter, _ as the shape parameter.
Some characteristics of a Iognormal distribution are:
mean = exp(_ + 0.5o 2)
median = exp(_)
mode = exp_ - o2)
, coefficient of variation = exp(a 2) - i
variance = mean2[exp(¢_ 2) - I]
., Cement hydration may be regarded as a process of subdivision of void space, or interstices
between anhydrous particles. At the early stages of hydration, the void spaces between anhydrous
particles are large. With proceeding hydration, hydration products bridge the particles and divide

3
the large voids between particles into smaller pores. This phenomenology suggests the lognormal
model of pore size distribution in cementitious materials to have a physical basis.
A mixture of two or more Iognormal distributions is defined by a compound density function:
=zfiP , zfi=l

where fi is the weighting factor of the ith Iognormal sub-distribution P(x, _ti, _i), _i and c i location

parameter and shape parameter of the i th sub-distribution respectively. If an attribute of a system


can be described by two or more Iognormal distributions, it is very likely that two or more
phenomena are occurring in that system. In the case of pore size distributions in cementitious
materials, this may indicate that different origins and formation mechanisms exist for pores in
different size ranges.
A statistical method has been developed to detect if there is a mixture of two lognormal
distributions and to iteratively estimate the parameters in the compound distribution. A similar
approach can be applied to a mixture of more than two lognormal distributions.
The pore size distribution used to demonstrate this method is obtained from a set of

cumulative probability data, P(x) vs. x as obtained by MIP. x is the pore diameter, P(x) is the
cumulative percentage of pore volume with respect to the total pore volume. The graphical
approach to obtain first degree estimates of parameters of a mixture of two or more lognormal
distributions first requires the transformation of x to log(x). Next, the quantfles of N(0,1)
corresponding to P(x) are determined by consulting a standard normal distribution tabulation [e.g.
see ref. 14]. If the cumulative percentage is 50%, the corresponding quantile is 0; if the cumulative
percentage is 84%, the corresponding quantile is 1.0; if the cumulative percentage is 16%, the
corresponding quantfle is the negative of the quantile for (100%-16%L or -1.0; etc. If a plot of log(x)
vs. the quantJJes is linear, x obeys a single Iognormal distribution. The final step in the analysis
is to determine the weighting factors, by locating the intersections between neighboring linear
segments. The goodness of fit can also be checked by comparing selected characteristics of
distribution. The median is often used as an important characteristic of a distribution, and it is
readily determined from the experimental data. Because the median = 5".(fi)(mediani), where
median i is the median of the ith sub-distribution, which is exp(_ti), the overall median can be
estimated from _. Another important characteristic is the inflection point on a plot of pore

volume vs pore radius. The inflection point is easily approximated on an experimental curve. The
inflection point coexists with the intersection between the first two linear segments. The
corresponding x value of the inflection point between the first and second sub-distributions is:
exp [(G2_I - aI112_ / (G2 "¢_1)]

In order to check the generality of fitting pore size distribution data to a compound
lognormal distribution, data from different sources have been examined. Ordinary cement paste,

4
blended cement paste, and mortar hydrated for various lengths of time have been examined. The
results demonstrate that it is reasonable to fit pore size distribution in cementitious materials to a
compound Iognormal distribution. Further, the graphical method used is adequate to provide
initial estimates of parameters in the compound distribution.
Pore size distributions in real materials must exhibit upper and lower bounds. This is the
physical basis underlying the model developed by Diamond and Dolch. The lower limit was
assumed to have little influence on the distribution curve and was neglected in the reduced
diameter expression. It is questionable whether this assumption is valid when small pores are
involved, i.e., when x approaches x1. If the lower limit is included, their model is equivalent to a

four-parameter Iognormal distribution, in which the variate X is confined to the range L < x < U,
and X = (X - L) / (U - X) is L(_ ¢). An attempt to desc_be pore sizes ranging from about 1 nm to about
10 fan by a single four-parameter lognormal distribution was not successful. Because the
simulated compound distribution is fitted to the pore size distribution determined by MIP, it
assumes 100% cumulative probability at the smallest pores intruded by mercury. In other words, it
assumes that the probability that the pore diameters are smaller than those intruded by mercury
is zero. This, of course, is not the real lower limit of pore size. If the real lower limit of pore size
and the percentage of the MIP porosity are both known, the pore size distribution over the range of
pore sizes determined by MIP can also be simulated by using the method we have developed.
The use of the distributions to describe porosity should have physical significance. The first
sub-distribution of the three may be regarded as describing the size distribution of coarse pores.
Pore sizes may extend to include voids. The third sub-distrlbution may be regarded as describing
the size distribution of fine pores. Pore sizes may extend to gel pores. The middle one represents
capillary pores.
It is the pores that belong to the sub-dlstribution representing the finest porosity that are
created by the hydration process. Because the majority of this porosity exists in the hydration
products that are forming, it is these pores that control the kinetics of hydration. Altematlvely,
from the viewpoint of permeabillty-pore structure or fracture mechanics-pore structure
relationships, the majority of porosity in this range is not important. With respect to
permeability, it is well recognized that only pores having diameters greater than some value
significantly contribute to permeability. We have demonstrated that an inflection point can be
calculated using the compound Iognormal model and based on our analysis.
Another characteristic that may be important to permeability is the mean square pore
diameter, or the second moment of pore diameter distribution. If one combines the classic Darcy's
• law and Poiseuflle's law, one can relate the permeability coefficient k to the mean square pore
diameter k = e <D2> / 32 where _ is the porosity and <D2> the mean square pore-diameter. This is
a classic model describing permeability of porous media, assuming that pores axe tubes which are
not interconnected. We have observed that a characteristic pore dimension and a tortuosity factor
are two indispensable variables in all sensible permeability models. If the pore dt_meter
distribution can be modeled by a compound lognormal distribution, the mean square pore
diameter can be readily determined using the equation: <D2> = Z fi exp[2 (iti + _i2)]. The inflection

point, or the mean square pore diameter, can be described in terms of a distribution expressed as
relative pore volumes, as relative pore numbers, or as relative pore surface areas. Depending on
the physical or mechanical property influenced by the distribution in porosity, it may be
beneficial to consider the pore structure in terms of pore surface areas or pore numbers. For
example, the Kozeny-Carman relationship relates permeability to porosity: K ~ e3/(1 - e)2/$2
where e is the total porosity and S is the total surface area of the pores.
This equation requires a value for the porosity expressed in terms of pore surface area. Thus,
multiplicative property of the Iognormal distribution, which allows the interconversion between

volumes and surface areas, anows, in turn, MIP pore volume data to be expressed in appropriate
terms. This attribute, coupled with the ability to deconvolute porosity data, suggests that a basis
has been identified which may allow a more fundamental understanding of relationships between
the behavior of cementitious materials and their pore structures.
In conclusion, a refined model for describing pore size distribution has been developed. This
in turn has been integrated into a model for the prediction of permeability. Finally, an
experimental apparatus to rapidly determine the permeabflities of specimens has been developed.
The predictions from the model show reasonable agreement with experimental values when used
to calculate permeability.

6
A Model for the Distribution of Pore Sizes In Cement Paste

O. Shi, P. W. Brown and S. Kurtz, The Materials Research Laboratory, The PennsylvaniaState
• University,UniversityPark, PA 16802

ABSTRACT

The pore size distributionin cement paste over the range of pore sizes interrogatedby high
pressuremercury intrusionporosimetrymay be describedby a mixtureof two lognormaldistributions.
The compounddistributiono( poresizes may be givenas:

)2 (l-f) [(log x-P.2) 2


p(x) = f exp- [(log X'lJ.1 exp - 2 ]
_r1 xV_-_'_ 20 2 ]+ a2x'V_'x 2(:12

where p(x) is the probabilitydensityfunctionof poresol sizex, f and (14) are the weightsof
sub-distributions,g.1and P2 are thelocationparametersofsub-distributions,
and 01 and (_2are the
shape parametersof sub-distributions.
These two sub-distributions
may representthe larger and
smallercapillaryporesrespectively.The changesin the sub-distribuUons
and the compound
distributionas functionsol curingage and water-to-cementratio"
are discussed.

INTRODUCTION

The abilitytowedicate the macro-properties


andbehaviorof porousmaterialsin generaland
those of cement-basedmaterialsin particulardependsonthe abilityto mathematicallydefinethe pore
structuresin these materials.In spiteof variety ol limitations,
mercuryintrusionporosimetry(MIP) is e
methodcommonlyusedfor determiningpore sizedistributionsof a broadrangeof pore sizes. It is
believedthatthe pore size range amenableto MIP analysisisthat whichaffectsmanymacro-
properties. It is of interesttodevelopa modelfor thedistribution
of pore sizesin cement pastes
determinedby MIP.
Previous investigationshave shownthatthe poresize distributionsin cement pastes
determinedby MIP couldbe describedby iognormaldistribution
providedthe pore diametersare
transformedto reduceddiametersby the equation:

Xr= (x- Xl)(Xu-Xl)/ (Xu- x)

where xr is the reduceddiameter,xu and xI are the upperand lower limitsofthe pore size range. The
lowerlimit was assumedto havelittleinfkJenceon the distributioncurve andwas neglectedin the
" reduced diameterexpression[1]. However. furtherwork.in whichhigherintrusionpressureshave
been applied,suggeststhata singleIognormaldistribution
may be inadequateto includethe smaller
pores. Additionally,the assumption
that the lower limitisnegligiblemay be invalidwhen smallerpores
are involved.
Based on the presentlyavailabledata, it appearsIhatthe pore size distributionin cement paste

7
overthe rangeofporesizesdetermined
byhighpmssuraMIPmaybedescril:xKI
by a mixtureof two
iognormaldistnl:utions.
Thesub-distributions
mayrepresent
thelargerandsmiler _ pores,
respectively.
Thevalidityof assuming
twolog-normally
distributed
poresizescanbetestedbya robust
statisticalmethod[2]. A program
basedonthismethodwasdeveloped
todetectwhethera poresize
_strroutionis a mbdure
oftwoiognormaldistributions.
Ifit isa mixture
oftwolognonnaldistributions,
threeparameters associated
witheachofthetwosub-distributions areestimated.Thethree
parametersforeachd_ion are:
theweightsofthelargerandsmallerporesizedistdoutions,
f and(1.q;
thelocatio
nparametemof
thelargerandsmallerporesizedistnbutions,
ILI andI_2;
theshapeparameters ofthelargerandsmallerporesizedistributions,
01 and02;
Thesizedistd0ution
ofporesincementpastesdetermined by MIPcanthenbedescribed
by a
compounddistribution
fromthesesixparameters.
Thechangesoftwosub-distnbutions,
withcuringageandwater-to-cement
ratio,aswellas that
of thecompound
sizedistnbution
areconsideredinthispaperasa preliminary
efforttounderstand
the
physicalmeanings
ofthesetwosulPdistritx#tions.
Itis a commonphysical
occurrence
thatphenomena
in a systemaredescribed
bythesumo(two
lognormaldistnbutions.
Thegeneralphysical
interpretation
isthattwoindependent
phenomena
can
governthatoccurrence.Inthepresentinstance,
thissuggeststhattheoriginsand themechanisms
controlling
the sizesoflargerandsmallercapiilar/poresmaybedifferent.Otherimp,cationsofthis
modelmayinclude
the possibility
ofdescrC_ng
thewholerangeofporesizesfromtheporesize
distnbutlons
determined
bydifferenttachniclues.

FUNDAMENTALS
OFTHELOGNORMAL
DISTRIBUTION

Thenormaldistribution
probability
densityfunction,
n(.u,o),is:

"X'=, I ex "(x'p')2'
Pt 2a'7 2
where p(x)isthe probability
thaithe randomvariable,x, willhavea valuex,p is themeanand a isthe
standarddeviation.
The mean,14,mayalsobe referredtoas thelocationparameterandthestandard
deviation,0, mayalsobereferredtoastheshapeparameter.
A variatex is iognormally
distributed
ifx - logy wherey isa positivevariate,(0 < y < ,,), whichis
normallydistributedwithmeanI_ and standarddeviationa. We thendefinethatx is Iognormally
distributed
andthe probability
densityfunction,I(_,a),is:

p(x)= 1 x-1_)2l
ox_/_z exp'[(l°g 2 °
20
wherex istherandomvariable,theporediameterinthiscase.Therangeofxis 0 < x < --. p,isdefined
asthelocationparameter,a is definedastheshapeparameter.Themean,medianandmodeareas
follows:

8
mean = exp(p.
+ 0.502)
median,exp(p.)
mode = exp(l_
-02")

Figure 1 Isthe comparisonbetweennormaldistributk)nn(O,O.S)and the Iognormaldistribution


I(O,O.S)[3]. The normaldistributionis symmetricabout the meanwhilethe IogrlormaldistdbuJionis
positivelyskewed. The normaldistributionadsesfroma theoryol elemanlaryerrorscombinedby a
ad_tive process,while the Iognormatdistributionarisesfrom a theoryof elementaryerrorscombinedby
a muItipUcative
process. The differencebetween Iognonnatand normaldistributionsas descriptorof
processis similarto thatbetween the geometricand the arithmeticmeans as measureof Ioca_n.
The derivationof the multiplicative
modeladsss fromviewinga posiUvevadate as a measureofa
discreterandomprocess.Atthe jth step, thechangein thevariateis a randompmpoflionof thevalue
of variateat the (j-1)th step. Thus,

xj- xj.1 = 8jxj.1

where xj isthe value of variateat the jth step, xj.1 the value at the (j-1)thstep, and 8] the smallrandom
proportionalgrowth/degradationratethattakes the processfromthe (j-1)th step tothe jth step.

0.8 I i I

Q6

0.4

0
-4 -2 0 z 2 4 6 8
<
w--Z
. 004
OLiJW
:S =S=S

Figure 1. A comparison between a normal distribution, n(O, 0.5), and a IognormaJdistirbution,I(0,


O.5).The mode,median, and mean of the IognormaldistnbulJonare shown.

g
It followsthat

xj = [['[(1 + 8i)]xo
Iog(xj)= Iog(xO)+ T..Jog(1
+ 8i) = Iog(xo)+ T._

where i = 0,1 ........ j. xo is the initialvalue of the vadate. Applyingthecentrallimit theoryto the sum of
p

the small random quantities,8i, resultsin log(x)havingan approximatenormaldistribution,i.e. x hasa


Iognormal distribution. Many physical and chemical processes, such as corrosion, diffusion,
pulvedzation,crushing,and cracking[4.7] may be approximatedby _ lognormal model.
Accordingto central limittheory, if log(x)obeys a normal distributionn(IJ.,a),then [Iog(x)-p.]/a
obeys a standard normal distribution,n(0,1). Thus, ifx1 obeys lognormaldistributionI(_,a), x2 obeys
standard normal distribution n(0,t), and if the cumulativeprobabilityp{x1 < x1,q} = p{x2 _;X2,q}, we
then have:

x2,q = [log(x1,q) - _]/o

In statistics,x 1,q and x2,q are defined the quantileof orderq of I(p,a) and quantileof order q of
n(0,t) respectively. Thismay be rewrittenas:

log(x1,q) = aX2,q+ I_
so that the locus of Ix2,q, log(x1,q]is a straightline. Inotherwords,it x1,q and X?,qare quantilesof the
same order q of I(p.,a)and n(0,t), then the locusof (x2,q, x1,q) is a straiglll lineon a semi-logscale.
Suppose now that the cumulativeprobabilityp(x2 < X2,q},insteadof the quantilex2,q, is ruledon an
axis with linear scale; this is the basis of logarithmic probabilitypaper. If plotting the cumulative
probabilityP{x < x} againstx on the logarithmicprobabilitypaper results in a straightline, we then
reasonably predict that the variate x obeys the lognormal distribution. This is one of the mostuseful
propertiesof the lognormaldistribution[3].

MIXTURE OF TWO LOGNORMALDISTRIBUTIONS

The mixtureof twolognormaldistributionis definedby a compounddensityfunction:

p(x = f exp - [(log X-lZ1) 2 ] + (l-f) exp"- [(log x'g2) 2 ]


_r1 x_/2zz 2 a 12 _r2 x _ 2(/22
where f is the w6!_ht of the first sub-distributionI(I.Zl,al), (l-f) is the weigh of the second
sub-distributionI(p2,a2) Pl and P2 are the locationparametersof two sub-distributions,repectively,
and <71and a 2 are the shape parameters of two sub-distributions,respectively. For the first
sub-distribution,the statisticalcharacteristicsare: ,,
mean 1 = exp(_.l + 0.5 o'12)
median 1 = exp(_)
model = exp(141 - a12)
For the second sub-lognormaldistribution,the counterparts are:

10
For the secondsub-lognormaldistribution,the counterpartsare:

mean 2 = exp(_2 + 0.5 (;2 2)


median 2 ,, exp(i.z2)
. mode 2 = exp(p.2 - <_22)
A statisticalmethodhas been presented to detectthe mixtureof two lognormatdistributions,

and to estimatethe valuesol f, P'I, P'2,(xl and (_2[2,7,8}.


A microprocessor-basedcomputer program has been developed at the Materials Research
Laboratory,Penn State University,to implementthis method. Withinthe context of the presenttopic,
this programfirst plotsthe experimental cumlative pore size distributioncurve on the log-probability
scale. If the curve appearsS-shaped, it is likelythat the distributionis a mixtureof two Iognormal
distributions. The program then selects a point, which may be but is not constrained to be the
inflectionpoint,as the startingpoint in the estimationof the value of l, thecorrespondingcumulative
probability.Thispointis also used to separate the wholedataset intotwosubsets. Foreach subsetol
data, the programdoes regressionanalysis on porediametersvs. quantilevaluesof n(0,1). The slope
and interceptof the regressionline are the valuesol p.and_. Fromthe estimatesoff, (t-f), P'I, P2, al
and (72.the compounddistributionand the cumulativeprobabilitiesare calculated. Then the program
comparestheexperimentalcumulativeprobabilityand the calculatedcumulativeprobability,and allows
the use ol iterativeprocedureuntil satisfactoryresultsareobtained.

EXPERIMENTALDATA

Foursetsof MIP dataolotainecl


from the[9,10]literaturehavebeenanalyzed. Sets 1 and 2 are

froma cementpastehydratedat a water-to-cementratioof 0.4 for one and28 days, respectb/ely.The


pore sizes rangelrom 2.9 nm to 700 nm. Sets 3 and 4 arefrom pasteshydratedfor 38 daysat
water-to-cementratios0.4 and 0.35, respectively.The rangeof pore sizesfor sets3 and 4 is 1.9 nm to
140 nm. Comparisonof calculated distributionswillbe madebetweenset 1 and set 2 to showthe effect
of age on pore structure,and between set 3 and set 4 to showeffectol water-to-cementratioon pore
structure.Forsimplicity,set 1 willbe referredto as: 1-daydata, set 2: 1-monthdata, set 3:0.4 w/c data,
and set 4:0.35 w/¢data.

RESULTS

In thefollowingfigures,"1"denotes the firstsub.distnbution


whichis the size distributionof the
largerpores,and "2"denotesthe secor_dsub.distribution,the size distributionof smallerpores.
Figure2a andb provideexamplesshowingthe estimationOfvaluesof f, 14and a fromthe MIP data. The
values ofp,and (; arethe slopesand interceptsof regressionlineswhichwereplottedon a natural
Ioganthmic-normal
quantilescale, not simply theslopesand interceptsof thestraightlines shownon
thatfigure. An interestingpointis that all the valuesoff are near0.9. Thispoint will be discussed
further.
Rgure3 comparesthe experimentalcumulativeprobabilitiesand thecalculatedcumulative
probabilities.SSE is the sumof square of errors,givenby the equation:

11
e.._=- ,. Experimental Data
**.m-I ---- 0.85
f- Regressions
>, **.=,- ol = 2.518 I_l = -0.647
:----
.Q
o2= 0.882 lz2= -0.131
0 90.Om- 1
.¢3
O t
L I
(3. t
50.01- I '

__o I ,,
%
lO.Om- t• %

b
0.10m- %
0 01s ....
1 10 100
Size (nm)

eo.om- Experimental Data


..u- --- Regressions
>, ...a=- f -- 0.90
_1 5.354 I_1= -1.055
•-- _z - 1.855 p2 = -0.237
.43 X
._0 90.Q=- X

Q. X _

_ 10.01-

0.1(_- a +
0.01l-- i

10 100
Size (nm) +
Figure2. Thepa_em f.rL.ando olXained
froma proba_lilyplolforpastehydrated
for(a)onedayat
wit ,, 0.4;Co)38daysat wit, 0.35.

12
N.t- N.IIIB=

W." t I:'xperimentol 04to N.m. _ ,, [xperimentol 04ta

0
W.m

re.D-
_.B°

,m.D.

tO.m- IOAI,
U 1,0_- t._.
._ .e.i. --- (stimoted__0ota w,-. _ --- [stimated 04ta
• ¢ ••_
e.lm_ iLim._ ,,
eom oow •

....... ,i siz.(n.,)'= si_;.(_) "


IQ.NB- N._-
w.'.,- \ (xperimentol 0oto w.m- I (xporimentol 0at0
\ --- (,timot.d Ooto I --- (,timotod 0_o
\
N.IB- IO.0l-

_ W.mo _,ll-

SO.I° . _ tO.m-
_w.m- _ Ill.m.
Lz_ - b • • • o - t_ • d ,•
O.Ota- . .... e.otl. . ....... m

,0 $iZ0 (rim),ol $iII (nm) ,m

Figure 3. A comparison of the exl=edmentai and calculated cumulative pmbabiUties based on the use
ot two lognonnal distributions to model the pore size distribution (a) one day at w/c ,, 0.4; Co)28 clays at
w/c - 0.4; (c) 38 days at wlc ,, 0.4; (d) 38 clays at w/c = 0.35.

SSE = I;(px - pc)2

where Px and PC are the expenrnental cumulative probabilities and calculated cumulative probabilities,
respectively. SSE can inclicato the goodness of fit.

Another way to indicato tho goodness of fit is by plotting the expodrnentat probabirdias against

the calculated wobabiities, which is actually the graphical express_n of SSE. Figure 4, which shows

the comparison for the MIP curve for 38 days at 0.4 w/c. is a typicalexample of this.

FigurO 5 shows tho caJculated SUb.diStributions, as well as the calculated compound


distributions, for the 4 data sets. It is worth noting that it is difficult to express the different weights of

_- sub-distributions graphically. Thereforo, we modified the calculated compound distribution to a mixture

of two sub- distributions with equal weights as follows:

i _ if

loo(_-)-(pl
.logf) 2 1 _g(_)-(.2-_(1-f)) 2
p(x). 1 exp-[. ] + exp-( " ]

13
1.00 -

0.80 -

0
i...
N
0.60 - /:
/,,,

0 0.4.0
..,_;
E
CL ...../..
×
0.20 - /

0.00 / I I I I I
0.00 0.20 0.40 0.60 0.80 1.00
Est. Cure. Prob.
Rgure 4. An exampleof the goodnessof fit shownthe samplehydrated38 days at w/c. 0.4.

Rgure 5 is thegraphicalexpressionof the modifiedcompounddistdbulionS.


Rgure 6a showsthe effect of increasinghydrationtime on the calculatedcompounddistribution.The
shiftof means and mediansof sub.distributions
is obvious,as are the twocrossoverpoints wherethe
sub-distril:uUonsintersect.

Figure61)shows the effecl of water-to-cementratio on the calculatedcompounddistribution.


The shiftsin the mean, median and crossoverpointare also easilyobserved.

DISCUSSIONAND IMPUCATIONS

ASmay be seen from Rgure 2, the f values (weightingfactors)am near 0.9 forthe distribution ._
ol the largerpore sizes. This suggeststhat, ifthe smallerpores areignored,the pore sizes mayfit a
singleIognormaldistribution. As mentionedearlier,previousinvestigationshave shownthatthe pore

14
O.OJ• &IS •

"_, lkN,
u Compounddistribution Compounddistribution
:_ --- Sub-distribution8 --- Sub-distributions
ta.

o.m .... -........ o._ _" > .........


'° sizsC,m)i_' ...... " s;'_o(,,,,) 'i_.
0.0,1 L_I 1

:_ .o
u
¢ Compounddistribution _L_S Compound distribution
_o.ol ------ Sub-distributions _
LI,. Sub-distributions

m ._LI@
c ¢
w •
0 0

i J °
o

OOI .- . ._ : ................. G, e.BI .................


10 100 tom t° 101
Size (nm) Size (nm)

Figure5. The probabirdy densityfunctionsof the calculatedcompounddlstnl=utlon


and o4the two
subdistributions
basedon the iognormalmodel(a) one day at wlc = 0.4; 28 daysatw/c = 0.4; (c) 38
daysat wlc = 0_4;(cl)38 claysat w/c= 0.35.

size distributionin cementpaste couldbe desatbed as a singleiognormaldistributionprovidedthat the


pore diametersaretransformedtoreduceddiameters[1]. By examiningthe MIP data used in that
previousinvestigation
(Fig. 6, and Table II in reference1), whichwas obtainedforpastes similarto
thoseanalyzedin thispaper(w/c ,, 0.4, age = 28-61 days), it was observedthatthe smallestpore
diameterused tofita singleiognormaldlstdbuttonis 10 rim. It seemsthatin on_r to includesmaller
pores, it iSnecessan/tOuSea mixlureof two Iognormaldistributions.FromFigureS,it was foundthat
the (_ametersofthe larger poresrange the orderof magnitude10 to 100 nm,whilethoseof thesmeller
pores are lessthen 10 rim. Resultsalso suggestthat the f value doesnot changeappmdablywith age
or water-to-camerara_o,in spiteof the shiftin the crossoverpoint. Thus,0.9 may be an appropriatef
value for the la_er pores.
Allreal panicleor poresize distribulionsmusthave some maximumand m_imum size. This
was the physicalinterpretationassociatedwith the use of reduceddiametersandw_ththe introduction
of the upperand lowerlimitson theparticleor poresize range[1,5,6]. Rgurs 5 showsthatthe
maximumsize does existwhere thecorrepondtngdensity functionapproacheszero, thoug_no
maximumsize is explk_tlydefined.On the otherhand, based on the cuwentlyavailalMeMIP data, it is
impossJl_eto locatethe real minumumsize. It might notbe validtoassume thatthe minimumsizeis _
negligibleif smallerpores, e.g.thosewithdiametersless than 10 nm, are included.Forthis mason
Figure5 leaves thesmallerporedistributioncurvesopen at the lowerend.

15
O.OSO

day old a
_.
-- 0.040 ---- month old

t_ Age: 1 day 28 day_


> 0.030
•_ Mean 1 386.64 76.17
cq3 Mean 2 6.58 4.21
ca 0.020 Median 1 211.37 57.81
>, Median 2 6.39 4.17
Crossover 11.52 6.18
.120.010 x
.Q I x
0 "_
L %

0.000
100.0 4OO.0
Size (nm)

0.200
I

c .... 0.35
.mo _
"co 0.150
:3
la.
l W/C: 0.40 0.35

>, Mean 1 17.46 15.28


"_ o.1oo Mean 2 2.75 2.44
e. Median 1 13.93 12.40
Median 2 2.44 2.41
>, Crossover "3.61 3.15
"--0050--
,mlll • _""_
r,l
0
r's
0
l,_
13.
0.000
0.0 20.0 40.0
Size (nm)
q:

Rgure 6. (a) The vafla_ons in the cak:ulated I_mba_fW density functions for Ihe pore size disffibutions
(a)betweenI and28daysatw/c- 0.4;(b)asalunclion
o_w/cratioataconstant
ageo138days.

16
It an a_bute of a systemcan be describedbytwo Iognormaldstdbutions,itis veryikely thattwo

phenomenaare __-__-rdng
in that system. In the presentinstance,this suggeststhat the odgtnsand
the formationmechanlsmeoflargerand miler capilla_ poresmay be differenLFigure6a showsthat
the largerporesize distdbutionshiftsto the leftmarkedlybetween 1 and28 days. The meanand

2
meo_anof the largerpore sizedistributiondecreasefrom387 nmand 211 nmto 76 rim and 58 nm,
respectively. On the otherhand,the decreasesin the meanand medianofsmallerpore size
_stnl:ution are relativelysmall The shape of the curve representingthe_ pores also changesa
great deal sugge_lng thatmanyof the largerporesdecreasesignificantly
in size. AlternaJvely,many
of the smallerporesare alreadypresent alteronlyone day of hydration.The obviousspeculationisthat
the smallerporesresulttramhydra_onwhilethe largerones are formedir_i_ly as intersticesbetween
anhydrousparticles.
Figure6b showsthat the sub-distdbutionsobserved at 38 daysof hydra_onexhibitmedian
pore sizes relatedto theirwater-to-cementratios,0.4 and 0.35. ComparisOn
w_ththe changes
observedin Rgurs 6a, suggeststhat the water-to-cemantratiodeterminesthe initialsub-distributions,
whilethe age determinestheevolutionof the sub-distrtbutlons.
However.more samplesneed to be
examinedto inveetigatethe pos___h4e
existence of the (lfferent originsandfonna_n mechanismsof
pores in cement pastes.
Porosityin cememrangesfrom millimetersto nanometersin cementand manydifferent
techniqueshave been usedto determinethe distributionsof pore sizesin these ranges[11,12].
Inevitably,these measureddlstdbulk)neoverlap. If the uncertaintiesassociatedw#h these
overlappingregionscouldbe solved, it may be possible to describethe entirerangeof pore sizesin
cement paste. The entiredistritx_ionof pore sizes in cement pastesfrommiimetersto nanometers
mightbe a mixtureof multiplelognormaldlstdbutions,eachsub-dlstdbulion
mpresentlngthe poresize
distdbutlondata determinedby one spodflc technique,and reflectingthe spod_ odginand formation
mechanismof poresin thatcorrespondingsize range.

CONCLUSIONS

1. A mixtureof two iogrlormaldistdbutioneappearsto deschbe poresize (IMdlxjlion detmned


by highpmesureMIP analysisand suggestsdifferentoriginsandfonnalionmechanismsof
poresin cement pastel.
2. The diametersof the larger poreersnge the orderof magnitude10 to 100 nm, wNle thoseof
the smallerporesare less than 10 nm. The sub<istrilxntonofla_11er
poreehas a 90% wei_t
in the _ sizedistntxltion.
3. The water-to-cementratioseems to determinetheinitialsub-distribution,whilethe aging
processatfeds the evolutionof pore structures.
4. Additionalsamples,includingthoseof blendedcements, needto be exanVned,to furmm
eluddete the existenceof the differentoriginsand formationmechanismsof poresin cement
pastel

17
ACKNOWLEDGEMENT & DISCLAIMER

The msearcJ1
describedhereinwas sul:fx)rtedbythe StrategicHighwayResearchProgram
(SHRP). SHRP is a unitof theNationalResearchCouncilthatwas authorizedby section128 of the
Sudace Transportationand UniformRelocationAssistanceAct of 1987.
This paper representsthe viewsof the author(s)only,not necessarilyreflectiveof the views of
the NationalResearchCouncil,theviewsof SHRP, or SHRP's sponsors. The resultsreported here are
not necessarilyin agreementwiththe resultsofotherSHRP research activities.They are reportedto
stimulatereview and discussionwithinthe researchcommunity.

REFERENCES

1. S. Diamondand W. Dolch,J. Colloid& InterfaceSd., _ 234-244 (1972).


2. E. Fowlkes,J. Am. Stat. Assoc., 74. 561-575 (1979).
3. J. AJtchisonand JJ_.C.Brown,The LoonormalDistribution.CambridgeUniv.Press,
Ca_ (1957).
4. P. Tobiasand D. Trindade,_ Van NostrandReinholdCo., New York, 1986.
5. R. Irani,J. Phys.Chem., 63. 1603 (1959).
6. R. Iraniand A. Callas,ParticleSize: Measurement.Intemmtation.andAootication,Wiley,New
York (1963).
7. D. Hosmer, Communicationsin Star., .1,,217-227 (1973).
8. V. Hasselblad,Technometrics,_ 431-444 (1966).
9. D. Shl, M.S. Thesis,Purdue University(1984).
10. A. Kumar,Ph.D. Thes_s,Penn. State University(1987).
11. R. Feldman,this Proceedings.
12. R. Gerhaml, this Proceedings.

18
LOGNORMAL SIMULATION OF PORE EVOLUTION DURING
CEMENT AND MORTAR HARDENING

Dexiang Shi, Weiping Ma and Paul W. Brown

Materials Rese_h Laboratory


The Pennsylvania StateUniversity
University Park, PA 16802

ABSTRACT

A model to describe the pore sizes in cement paste and mortar, as determined by hi.gh pressure
mercury intrusion porosimetry, has been developed. The model describes porosRy using a
compound lognormal distribution. For given material under a given set of curing conditions, the
weighing factors and shape parameters of two sub-distributions in the lognormal model may be
considered as constants, while the location parameters may be related to curing time and the
relationship can be quantified. Therefore, it is possible to predict both the pore size distribution in
cement and mortarat any age as well as the evolution in pore size during curing.

INTRODUCTION

Cementidous materials are being used in the immobilization of radioactive waste. Conventional
concrete structures are being used as engineered barriers. Low level liquid wastes are being
immobilized in cement-based grouts. While these applications differ from those involving normal
concrete structures,the porosities of cernentitions materialsare important in determining the ability
of a structure to meet its functional requirements, virtuallyregardless of the specific applications. It
is the distribution of porosity throughout cement matrix and at cement-aggregate interfacial zones
that, in concert with environmental variables, determines the macroscopic rate of transport of
species.

The development of an adequate description of thepore size distributions in cements has recently
been accomplished [1]. Our work has shown that the pore size distributions in cement pastes over
the size range determined by the high pressure mercury intrusion porosimetry (MIP) may be
described by a mixture of two lognormal distributions [1]. The present paperextends the model to
cementitious systems containing aggregate.

Many physical and chemical processes may be approximatedby the lognormal model [2-5]. It is a
common physical occurrence that a system can often described by a mixture of two lognormal
distributions, which .govern two different phenomena occurring in that system [6]. Cement
hydration may be reganted as a process of subdivision of void space, or interstices between
anhydrous particles [1,5]. In cement paste and mortar, the origins and the mechanisms controlling
the sizes of largerand smaller capillary pores may be different, suggesting the Iognormal model of
poresizedistribution incementpasteandmortar tohaveaphysical basis.Previous investigations
ofporeevolution incementpastehaveshownthat thedistributionshiftstothesmaller poresize
withincreasing age,andthatthemediandiameter andthreshold diameter ofporesdecreases with
increase age{7]. Thispaperwillshow that, fora givensetofcuringconditions, theweighing
factors, fand (I-0,andshapeparameters. Ol ando2,ofsub-distributions may be considered
constant, while the location parameters, ot and o2, of sub-distributions may be related to curing
time. Thus, it becomes possible to predict the evolution in pore size during the curing of cement
and mortar and to predict pore size distributions of cemem and mortar at any age.
BACKGROUND OF LOGNORMAL DISTRIBUTION

A variate X is lognormally distributed with the location parameter tt and the shape parameter0', if
• X = logY where Y is a positive variate, which is normally distributed with mean tt and standasd
deviation o. A compound lognormal distribution of pore sizes can be expressed as follows:

19
where p(x) is the probability density function for pores of size x, f and (I - f) are the weighing
factors of sub-distributions" l.tl and _t2 are the location parameters for sub-distributions" o! and 02
are the shape parameters for sub-distributions. Based on a robust statistical method [8], a program
has been developed to detect the existence of a mixture of two lognormal distributions. For
cementitious materials, the two sub-distributions may represent the larger and smaller capillary
pores, respectively [I].

EXPERIMENTAL

Cement pastes were prepared having water-to-cement ratios (w/c) of 0.3 and 0.5. These were
cured for I, 3, 7, 14 and 28 days and for I, 3 and 7 days, respectively. Pore size measurements
were carried out at these ages using mercury intrusion porosimelry (MIP). The details of MIP can
be found in reference [6]. Mortar samples were prepared at a water-to-cement ratio of 0.47. The
cement-to-sand ratio used was 0.623. Saturated surface dry (SSD) sand was used. Samples were
tested after curing for 1, 3, 7 and 14 day_. Triplicate MIP measurements were performed for each
sample.

RESULTS AND DISCUSSION

A Comnound Lolnormal Distributionof Pore Sizes in Mortar

Replicate measurements of pore size dismbutions in mortarshowed excellent agreement. Figure I


showsthe variauonamong thr_measunmmmtsofthe cumulative poresizedistributionsina mortar
sample cured for 3 days.

To evaluate the parameters of the compound lognormal diswibution, log x is plotted versus (log x -
p)/o • log x is defined the quantile of the mixture distribution and (log x - p)/o the quantile of
standard normal distribution [8]. Figure 2 shows Q-Q plots for mortar samples cured for 1, 3, 7
and 14 days. The slopes of the two roughly linear segments are initial estimates of Ol and 02, the
intercepts of the two segments are initial estimates of $tl and _2, respectively. The initial estimate
for f is the deflection point. These estimates are then iterated. Mathematically, the iterationcan start
using any one of these five estimates. They may result in different sets of estimated parameters,
though they may generate the estimated distributions with the same goodness of fiL Therefore, it is
necessary to establish the physical meanings and the reasonable ranges for the five parameters so
that the estimation is not only judged by the goodness of fit, but also by the predetermined,
physical meanings and ranges, The meaning of the weighing factors, f and (l-f), is evident. It is
necessary to establish the direct characteristics associated with the shape parameters and the
location panmctas.

Close examination of Figure 2 shows the linear segments at each end of curves to be almost
parallel to each other. This suggests that the values for the shape parameters, Ol and 02 in the
lognormal distributions of pore sizes are the same. Figure 3 shows the cumulative probability
distributions for the mortars cured for 1, 3, 7 and 14 days. These curves exhibit the same shape
behavior, which can be characterized by the slopes of the linear segments in Figure 2. Thus, the
shape parameters in the compound lognormal distributions may be considered inde.pendent of
curing time. Itfollows only the location pamn_ters alone may be related to curing un_ while
that
the shape parameters are related to other factors than curing time. Table 1 shows the estimated
parameters. Figures 4a-d compare the experiment_y obtained pore size distributions and those
calculated using the estimated paramem's. The mean of sum of squaredenms between the two are
0.001, 0.001, 0.001 and0.002, respectively, indicating very good fits.

2O
"_ _ Do,/

X,,v,

"6

o.

t)_ameter (rim)'Q° -°_o -,_o -,:mQuantile¢60


of ,.6ON(0.1)_6o
_b ,.6o
Figure 1. Reproducibility of porosity measure- Figure 2. Q-Q plots for mortarspecimens
meritson mortarspecimens cun_ for 3 days. cured for l, 3, 7, and 14 days.

Estimates of parameters in log-


normalmodelforporesizedisn'ibufion n

o--o ,,,
1 3.67 1.05-1.005 -0.16 0.87 _o
3 3.10 1.03 -I.0 -0.I6 0.85 _ T40ays_
7 2.80 0.96-I.0 -0.15 0.85 _,°
14 2.43 0.85-0.98 -0.14 0.84
8
c_, ....... ;b ...............
Diometer (nm_°
Figure3. Cumulativeprobabilitydiscibudonsfor
mo_ cm'edfor 1, 3. 7, and 14days.
Prediction of Pore Evolution Durine Cement and MortarHardenine

From above analysis, it was found that p! and P.2may be related to curing time. while the f, at
and o2 may be assumed to remain constant. Figure 5a shows that Pl for mortars cured for 1, 3
and 14 days (can be related to cta'ing time as in days, t).

P.l = 3.67403 t('°'ls63'tg) [2]

Figu_ 5b shows that g2 for moru_ cured for I, 3 and I4 days (can be related to curing time):

tt2 = 1.07456 exp (-0.016661t) [3]

It must be noted that equations 2 and 3 fit the limited, presently available data. Unlike in the use of
Iogm,,mal model to describe pore size distribution, there is no physical basis for these equations.

The 7-day dam were not used to obtain the above expressions. Rather, seven day values were
predicted from the above age-p relationships. The predicted values are Pl = 2.71 and P2 = 0.96,
assuming f = 0.85, Ol = -1.0 and 02 = -0.15. Figure 6 shows the comparison between the

21
experimentally obtained pore size dislribudon and that using the predicted parameters. The mean of
sum of.squared em3_ between the two is 0.0011 indicating a very good prediction.

• 1 Day
_'_ --- Estimated
._ _ %,,, ---Estimated
0 O "

"_o
0

Diameter (,nm) Diameter (nm)

7 Days 14 Days

-,% ___ :__ ,'_ ___ Estimated


_,_. Experimental "_ Experimental

_,_o
0

o! tO I_ o! I0 IW
Diameter (nm) Diameter (nm)

Rg,_ 4. Compa.,'iso. between expedmcm,_lly ob_dned CRg,_ 3) and caJcu_:d ix)n: size
disu'ibudons usmg d¢ iognormai model.

t_.• II.IIB--
_.m- Experimental

(t ** -.156349) * 3.67403 _ M_e=-

110.(11-
--- Predicted
n

O
_ ItLQI-
| •
- • t) • .0745
Ot_

l,o s.6oAge (day) '°'_° ,s.bo o.0.- . .....Size i_lom) . ..

Figure 5. Variations in I.tl and 1_2with curing. Figure 6. Comparison between experimentally
obtained and calculated distribuuons.

22
Figure7 showsthecumulative probability
distributions
forcementpastes
fabricated
atawater-to-
cementratio0.3andcuredforI,3,7,14 and28daysandthosefor0.5w/cratio cementpastes
curedforI,3 and7 days.The shapesofthecurvesinthefigures appeartodependsu'ongly
on
water-to-cementratio
butverylittleon age.Figures8 showsthecorresponding
Q-Q plots.
Ata
givenwater-to-cementratio,
thelinearsegmentsateachendappear parallel
toeachother.
Taken
' together,Figures7 and 8 suggestthatitshouldalsobe reasonabletoassumeconstantshape
parametersforcementpastesatdifferent
ages,giventhesamecomposition
andcuringconditions.

"'-..,,/c=o.s
O_ I _'k_. _ • X
.O114 "4& t',_,

°'.I
.Io.O.3 ,,,,
','-; ,/e
3 -- ",,,,

ol ll_iorneter (n Quontile of N(O,1)

Figure 7. Cumulativepore sizedistributionsfor Figure8. Q-Q plotsfor cement pastespecimens


cementpastes:w/c = 0.3 at l, 3, 7, 14, 28 days; havingpore sizedismbutionsshownin
w/c = 0.5 at 1, 3, 7 days. Figure 7.

Comparing Figures 7 and 8, the 0.3 w/c ratio cement pastesshowlitde change in pore size
distribution with curing time. Comparedto pore size dismbutionsfor 0.5 w/c ratio cementpastes,
there is significantly less large pore space available for size refinement. These data serve to
reinforce the idea that the water-to.cement ratio determines the pore structure during the setting
process and tha_it is from this pore structure that the pore evolution occurs [1].

Because the cumulative distributions of the 0.5 w/c ratio paste show noticeable shifts while
maintaining the same shape, the method described above was used to predict the pore size
diswibution in 3 day old cement paste. It was determined that f= 0.91, Ol - -1.1 and o2 - -0.24.
Figure 9 shows the age-stl and age-st2 relationships. From these relationships, the values predicted
for Stl and $*2are 4.19 and 1.15, respectively. Using the above values for the five parameters, the
predicted compound lognormal distribution of pore sizes in 3 day old, 0.5 w/c ratio cement paste
was obtained and shown in Figure 10. In this instance the mean of sum of squared errors between
the experimental and the estimated data is 0.0015.
SUMMARY

Based on presently available data, interpolation has been used to obtain It.values. In order to
predict the long term changes in pore s_ucture, it will be necessaryto obtain sufficient pore size
disn-ibudonsdata to allow exlxapolation. In addition, the relationshipsbetween w/c ratio and the o
and St values, and those between curing conditions and o and St values have not been fully
elucidated. Further acsessment of the effects of w/c ratio andcuring conditions on the parameters
in the compound Iognormal model are needed. However, in spite of these limitations, the present
work has shown that it is reasonable to assume constant values for o and f for both cement pastes
and mortars to establishment of simple relationships between curing time and St values. Given
sufficient data,it should be possible to predict pore evolution duringcement and mortar hardening,
and to predict the pore size distributionof cement and mortarat any age.

23
ImJllS,-
(t ,* -.0335962) .4.35424 0tern- * Experlmentol
_l .. --- Predicted

1tLm*
•, (t ** .0712061) * 1.24833 E "_
, ::1
_. 0 I._1- *

O.lm-
0.01e-

3, AQe (day) '_ize (rim)

Figure 9. Variations in Pl and £t2with curing. Figure I0. Comparison between experimentally
obtainedandcalculateddistributions.

CONCLUSIONS

I. A compound lognormal distribution model can be used to describe the pore size distribution
in themortarandprovidesthebasisforthe predictionof theporesizedistributionat any age.

2. It is reasonable to treatweighing factors and shapeparameters in compound lognormaIpore


sizedistributionsfor cementandmortaras constants.

3. Thelocation
parameters
inthemodel,_ canberelated timeforcementand mortar.
tocuring

REFERENCES

I. D. Sift, P.Brown andS.Kurtz, MRS Meeting, Boston,1988,(inpress).


2. I. AJtchison and I. A. C. Brown, The Lognormal Distribution, CambridgeUniv. Press,
Cambridge(1957).
3. R. kani, J. Phys. Chem.,63. 1603 (1959).
4. R. Irani and A. CaLlas,Panicles Size: Measurement. Internretation.andApnlicafion_ Wiley,
New York (1963).
, 5. S. Diamond and W. Dolch, J. Colloid and Interface Sci., 38. 234-244 (1972).
6. S. Kunz, S. Levinsonand D. Shi, J.Am. Ccram: Soc., (in press).
7. D. W'mslow, J. Mat., 5, 56*.585 .(1970).
8. E. Fowikes, J. Am. Star. Assoc., 74. 561-575 (1979).

ACKNOWLEDGEMENT AND DISCLAIMER

The research describedherein was suppo,ned by the Strategic Highway Research Program
(SHRP). SHRP is a unitof the National Research Council that was authorized by secdon 128 of
the Surface Tnmspo_tion and Uniform Relocation Assistance Act of 1987.
This paperrepresents theviews of the author(s) only, not necessarily reflective of the views of
the National Research Council, the views of SHRP, or SHRFs sponsors. The results reported
here are not necessarilyin agreement with the resultsof other SHRP researchactivities. They are
repottedto stimulate reviewanddiscussionwithin the research community.
c

24
CONCRETE MICROSTRUCTURE AND ITS RELATIONSHIPS TO PORE
STRUCTURE, PERMF.ABILITY, AND GENERAL DURABILITY

D.M. ROY, D. SHI, B.E. SCI-IF_ETZ,AND P.W. BROWN

Introduction

The durability of concrete is frequently associated with the transport


of dissolved species. Such transport may be considered in terms of
permeability. It is well recognized that transport occurs through a
continuous network of pores, which exist in the cementitious matrix of
concrete, as well as through the porosity which exists in the interracial
regions with aggregate. Unfortunately, however, the relationships between
concrete permeability and the pore structures through which transport
occurs are, at best, qualitative. It is the objective of this paper to describe
work leading to rapid and accurate measurement of concrete permeability
and the development of models to describe permeability of concrete in
terms of pore structure.

Poised Perm¢obility

Permeability measurements have historically been made with water


or various gases including oxygen, nitrogen, argon and air. Darcyian flow
has been used as the theoretical basis for the description of the observed
flow conditions. This mathematical treatment required information on the
sample dimension (cross-sectional area and permeation length), flow
properties of the fluid, a measurement of the flow rates and an established
pressure gradient across the test specimen. The vast majority of
experiments have been conducted at low AP in accordance_ with Darcy's
derivation and usually with ambient pressure being the lower bound on
AP. Experimental designs of this type are limited to values of permeability
that are typically to the nano-darcy range. In using apparatus typical this
type, water which has permeated through the test specimen is collected on
an LVDT, the spring constant of which has been matched to the mass of
water that would be collected in a reasonable laboratory experiment [Goto
and Roy (1981)]. The resulting displacement of the LVDT by the
permeated fluid is monitored and an equilibrium flow rate is determined.
The permeability is calculated from this flow rate. One limitation to this
particular experimental design is the evaporation of the permeated water
from the container which is affixed to the LVDT. A second limitation is
that the lower limit of operation for this apparatus is estimated to be 10-8
darcy.

25
To develop porosity-permeability models, measurements on
specimens having low permeabilities need to be carried out. To address
this need, the pulsed permeability approach was implemented.
Transient decays in pressure have been used in the pressure-pulse
permeability cell to successfully measure hydraulic properties of low
permeability materials such as cored samples of rocks and sandstones
(Brace et al. [1968], Hsieh et al. [1981], Neuzil et al. [1981]). To date, there
has been little work devoted to the application of this technique to
cementitious materials (Hooton and Wakeley [1989]), which generally have
higher compressibilities and permeabilities. In the transient pressure
pulse method, a jacketed sample is confined between two pressurized
reservoirs which contain the penetrating fluid. The pores of the sample
are also filled with the fluid. A confining pressure higher than the
reservoir pressures is maintained on the jacket. When the experiment
begins, the pressure in one of the reservoirs is suddenly changed to a
higher or lower value and the resultant pressure changes on the high or
low pressure side are measured as a function of time.
The optimal way of collecting data is to simultaneously measure the
pressure change in both reservoirs. Pommersheim and Scheetz [1989]
have recently discussed the advantages of this technique. Hooton and
Wakeley [1989] have emphasized the sensitivity of test results to
environmental variables when measuring water permeabilities of concrete.
Figure 1 presents a schematic diagram showing the experimental
configuration for the transient pressure pulse test and Figure 2 a detailed,
exploded drawing of the cell design. Before the test begins the entire
system is equilibrated to a constant pressure P0. Then the pressure
downstream reservoir (Pd) is suddenly deccreased to P1, while the
upstream reservoir (Pu) remains at P0. Pu will decrease and Pd will rise as
fluid is transferred between reservoirs. A constant confining pressure (Pc)
is kept outside the sample. By maintaining this pressure at a level
considerably higher than P1, leaks are prevented. However, Pe is limited
to avoid creep or micro-cracking.
Figure 3 is a schematic representation of how the upstream and
downstream pressures change with time for two typical classes of
materials. The figure is illustrated for the case where the two reservoir
volumes, Vu and Vd, are equal. Nomenclature is provided beneath the
figure. The curves marked A. are the pressure decay and pressure rise
curves for "tight rocks" such as granite which typically have low
permeabilities and compressibilities. Here the response of the two curves
is symmetric around a horizontal line drawn to the final pressure Pf
reached by both reservoirs. These data collected for an actual sample are
presented in Figure 4. As discussed by Pommersheim and Scheetz (1989),
the analysis of Brace et. al. [1968] predicts that this pressure will lie
midway between P0 and P1, while the analysis of Hsieh et. al. [1981] °
predicts that Pf will lie more towards the upstream side. The difference is
attributable to the fact that Brace and co-workers assumed in their

26
development that the compressive storage of the sample was negligible,
whereas Hsieh and co-workers did not. Pf can also be estimated from a
mass balance knowing the pore volume of the sample and the volumes and
initial pressures in the reservoirs (Trimmer [1981]).

Model Development

Previous mathematical models for pressure pulse testing have been


developed by Brace et. al. [1968], Lin [1977], Hsieh et. al. [1981] and
Pommersheim and Scheetz [1989].
As a model system consider the cylindrical sample depicted in Figure
1, having total volume V = AL, where A is the cross-sectional area of the
specimen and L is its length. The sample is confined between the two
pressurized reservoirs, the upstream one at Pu and the downstream
pressures are P1 and Po, respectively.
The partial differential equation which governs pressure changes as
a function of distance and time P(x,t) within the sample is given by:

o_p ,( &p ,_2 fl' # 6P (1)


6)(2 + #\6-x-) - k at

where P = pressure within the sample;

g = pore fluid viscosity;

k = sample permeability;

x = axial distance within the sample, measured from the upstream


reservoir;

t = time;

13' = lumped compressibility.

Equation (1) represents how the pressure P(x,t) varies within the
sample as a function of position x and time t. It is a non-linear partial
differential equation subject to one time and two boundary conditions.
These are given by:

P(x,0) = P0 initial condition

P (0,t) = Pu(t) boundary conditions

P(L,t) = Pd(t)

27
where L is the sample thickness, and Pu and Pd are the upstream and
downstream pressures, respectively.
Assumptions involved in the derivation of equation (1) include: one
dimensional mass transfer with constant transport area, constant
temperature, and constant physical properties (g, k, [_' and porosity).
These assumptions are the same as those stated or implied by previous
workers [Brace et. al. (1968), Hsieh et. al. (1981)]. Most of them are likely
to be met in laboratory tests with homogeneous corings. Sample
cempressibilities will he most likely to remain constant when the ratio of
the confining pressure to the initial pressui'e difference, i.e., Pc/ (P1-P0)is
high [Hooton and Wakeley (1989)].
13' is a lumped compressibility. It depends on the compressibility of
the fluid, 13, the compressibility of the sample, 13s, the effective
compressibility of the jacketed sample 13e, and the sample porosity e,
according to:

13' = (13e - 13s) + e(13 - [3e) (2)

A similar equation has been presented by Brace et. al. [1968] and Hsieh et.
al. [1981]. Of all three quantities be is the one which is least likely to be
known and which is also potentially the largest, especially in experimental
configurations where the sample is retained in a flexible rubber or plastic
sleeve. In effect this makes 13e, and thus 13', an arbitrary parameter. Neuzil
et. al. [1981] found values for this parameter which were several orders of
magnitude greater than the fluid compressibility.
By introducing the dimensionless distance z = x/L, time 0 = t/T, and
pressure P = (P- P0)/AP, equation (1) and its attendant conditions
become:

6z2 +
a,AdSP/
_,oz ,, = 6o (3)

(I) (II) (III)


where AP = P1 P0

with conditions: p(0,q) = Pu(q) reduced boundary conditions

p(1, q) = Pd(q) reduced initial conditions

p(z,0) = 0
¢

T = 13'ttL2/k is a characteristic time for the transfer of mass from the


high pressure to the low pressure side.

28
The terms in equation (3) have been labeled (I), (II) and (III). Term
(I) represents the transfer of mass between reservoirs caused by the
pressure difference. It is present in all formulations of the problem. The
unsteady state term (III) corresponds to the accumulation of mass within
the sample. If the observed experimental times t are much greater than
the characteristic time T, i.e., t >> T, then the accumulation term can be
neglected. This is a requirement of a steady state process. However, since
, the pressures at the two ends of the sample are continuously changing,
pressures within the system slowly adjust to accommodate these changes,
in effect adjusting to each new steady-state. Such behavior is called quasi-
static and the system is said to be at quasi-steady state.
Term (II) in equation (3) represents the dynamic response to the
compression of the jacketed sample. This non-linear term will be small
when the dimensionless compressibility 13'AP is small. Calculations show
that term (II) would be small if _' is of comparable magnitude to the
compressibility of the fluid, but, as discussed, this is often not the case in
practice.
Application of these equations to data from Figure 4 should result in
a strength line when stated at time vs.

Pu - PL
k

From the slope of this line the characteristic chance time can be
determined and subsequently the permeability. Figure 5 represents the
processed data and the regression fit to the data.
Pore Structure

Recent studies by Shi et al. [Shi et al., 1989, 1990a, 1990b] have
shown that the pore size distributions in cementitious materials may be
described in terms of a mixture of lognormal distributions. A variate X is
lognormally distributed if Y = logX where X is a positive variate, (0 < X < **),
which is normally distributed with mean _t and standard deviation ¢_ with
regard to Y. The probability density function of a lognormal distribution is
[Aitchison and Brown, (1957)],

p(x) = (2xa2) -1/2 X-1 exp[-0.5{(log(x) - lX)/o}2]

The range of x is 0 < x < _t. Ix is defined as the location parameter, a is


defined as the shape parameter. The mean, median, mode and variance
, are as follows:
mean = exp(_t + 0.502)
median = exp(lx)_
" mode = exp(].t - 0 2)
variance = mean2[exp(o2)-l]

29
The mixture of two or more lognormal distributions is defined by a
compound density function:

P(x) = _: fi p(x, _ti, a i)

Zfi=l

where fi is the weightingfactorof the ithsub-distribution


p(x,IA b cri),
and
Ix
i and c_i are locationparameterand shape parameterof the ith sub-
distribution,
respectively.A statisticalmethod was developedto detectif
thereis a mixtureof two lognormaldistributions and iterativelyestimate
the parametersin the compound distribution[Fowlkes,(1979)]. Without a
computer program, the iterative estimationof parametersis extremely
difficult.
However, one can use a simplegraphical method to obtainfirst
degree estimatesof parametersfor a compound distribution of two or
more lognormaldistributions.
The method is describedas follows.Suppose
that one has obtaineda set of cumulativeprobability data,P(X) vs. X
where X is thepore sizeand P(X) is the relative
percentageof pore volume
with respectto the totalpore volume. The graphicalapproach to obtain
estimates of parameters of a mixture of two or more lognormal
distributions
can be summarizedas follows.

Step 1: Transform X to log(X).

Step 2: Find the quantiles of N(0,1) corresponding to P(X), by consulting the


standard normal distribution tabulation which can be found in any
statistics text book. For example, if the cumulative percentage is 50%, the
corresponding quantile is 0; if the cumulative percentage is 84%, the
corresponding quantile is 1.0; etc.

Step 3: Plot log(X) vs. the quantiles.

If one single straight line is found, one can conclude that X obeys a
lognormal distribution. If more than a single straight line is observed, it is
likely that the porosity can be described by a mixture of lognormal
distributions.

Step 4: Determine the intercepts and slopes of each linear portion. These
correspond to _ti, and ci, i =l,...,n, respectively. From the intersections
between any two neighboring linear segments, one can find the weighting
factors, fi-

Step 5: Determine the quality of the fit by drawing lognormal distributions


using the parameters obtained and comparing them to the data.

30
An important characteristic of a mixture of lognormal distributions is
the inflection point. This is located on the experimental curve at the
intersection between the first two linear segments. The corresponding
log(X) value is:
4

[a2 lXl - al Ix2] / [a2 - all


7

Comparing the calculated inflection point with that from the experimental
data allows a comparison of the model fit.

Pore Structure, Percolation and Permeability

Among the models relating pore structure and permeability,


percolation theory-based models are physically the most meaningful.
Percolation theory assumes a lattice containing a very large number of
lattice points [Stauffer, (1985)]. It is assumed that a point can be
randomly occupied or vacant independent of the states of the neighboring
sites. However, depending on the situation, additional assumptions can be
made which influence site occupancy. Accordingly, occupied sites may be
isolated from each other or connected with neighboring sites to form a
cluster. When only a small fraction of the total lattice sites are occupied,
the probability, p, that a given site will be isolated is high. With increasing
site occupancy, the probability of the formation of a large cluster, which
extends from one side of the lattice to the other, becomes non-zero. Such
a large cluster is called "infinite-path." It percolates through the lattice in
the similar way that fluid percolates through porous materials along the
network of "infinite-path" pores. There is a statistically-based critical
point (percolation threshold), Pc, below which no 'infinite-path' can form.
It is clear that these concepts are relevant to the transport of liquids
through the "open" porosity in concrete. In the determination of
permeability, only the liquid which flows through the network of
connected pores is determined. The network of connected pores is
equivalent to the "infinite-path" in percolation theory. Therefore, it is the
structure of the infinite network, not the total porosity and its distribution,
which controls permeability. It is generally accepted that only pores with
diameter greater than some value effectively contribute to permeability.
The observed inflection point-on the cumulative pore size distribution
determined by mercury porosimetry locates the minimum diameter of
pores which are continuous through all regions of the hydrated cement
paste [Nyame and Illston, (1980); Mehta and Manmohan, (1980); Hughes,
(1985); Goto and Roy, (1981)]. It is this observation which provides the
connection between the lognormal model for pore structure, discussed
-. above, percolation theory and permeability.

31
A Preliminary PCrqol_tion-Permeability Model

Most of the percolation models which have been developed have


been applied to describing flow in sedimentary rock. This work suggests
that a power law may exist between permeability and a characteristic pore
dimension, and that a tortuosity factor which may be quantified by using
percolation concepts. These are important aspects of modeling pore
structure-permeability relationship in cement and concrete. "

In percolation theory, the relationship: f = C(p' - pc) n can be used to


relate f, the fraction of connected pores in the infinite network, to the
critical porosity, Pc. C and n are constants, Pc the critical porosity below
which there is no flow permeating at all, p' the probability that two pores
interconnect, and f is related to tortuosity [Guegeun and Dienes, (1989)]. It
has been suggested that n = 1.9 or 2 [Fisch and Harris, (1978); Berman, et
al, (1986)].
In the following, the three remaining parameters are calculated.
Assuming p' to be proportional to porosity p and represented by Pc', we
have f = C'(p- pc') 2, where pc'= Pc/c'. Pc' can then be obtained by
extrapolating the porosity-permeability curve to the point where the
permeability equals zero. C' can be obtained by fitting the experimental
data to the model. From very limited data on permeability and pore size
distribution, we calculated P.C' = 0.15 and C' = 0.23.
We chose the second moment of the pore diameter distribution,<D2>,
as the characteristic pore dimension. The predicted permeability is (the
derivation is omitted here):

k = (f/32) p<D2>

The second moment of a lognorrnal distribution is exp(2ix + 2a2).


In our compound lognormal model for pore size distribution,

<D2> = Y_fi exp(2 (IX+ a2))

This approach can be demonstrated using the data shown in Figure 6,


which are for a fly ash blended cement (fie = 0.536), mixed at a water-to-
solids ratio of 0.55 and hydrated for 28 days at 23°C. The value of<D2> was
calculated as 2383 nm2. The porosity and permeability were determined
to be 0.45 by MIP and 3.86 x 10-5 Darcy by flow-through permeability,
respectively. The calculated value of the permeability is 1.93 x 10-5 Darcy.
Although the prediction is satisfactory in this case, the model must be
validated by comparison to additional data to obtain reliable values for C',
Pc, and to verify its general validity.

32
S_mmary

We have briefly discussed the means for the rapid measurement of


permeability, the means by which pore size data obtained from mercury
, intrusion porosimetry can be quantified, the basis for using percolation
theory to treat the pore data, and the development of a preliminary model
to describe permeability.

33
References

J. Aitchison and J.A.C. Brown, The Lognormal Distribution, Ch. 2_ Cambridge


Univ. Press, Cambridge (1957).
P

D. Berman et al., Conductances of filled two-dimensional networks, Phys.


Rev. B, 3_, 4301, 1986.

W.F. Brace, J.B. Walsh and W.J. Frangos, Permeability of granite under high
pressure, J. Geophys. Res. 7_ (6), 2225-2236 (1968).

R. Fisch and A. Harris, Critical behavior of random resistor networks near


the percolation threshold, Phys. Rev. B, 18, 416, 1978.

E. Fowlkes, J. Am. Stat. Assoc., 74, 561-575 (1979).

S. Goto and D. Roy, The effect of w/c ratio and curing temperature on the
permeability of hardened cement paste, Cem. Cone. Res., 11, 575, 1981.

Y. Gueguen and J. Dienes, Transport properties of rocks from statistics and


percolation, Math. Geology, 21, 1, 1989.

J.D. Hooton and J.D. Wakeley, Influence of test conditions on the water
permeability of concrete in a triaxial cell, Pore Structure and
Permeability of Cementitious Materials, 157-64, L.R. Roberts and J.P.
Skalny, Eds., MRS (1989).

P.A. Hsieh, J.V. Tracy, C.E. Neuzil, J.D. Bredehoeft and S.E. Silliman, A
transient laboratory method for determining the hydraulic properties
of 'tight' rocks--I. Theory, Intl. I. Rock Mech. Min. Sci. & Geomech.
Abstr. 18, 245-252 (1981).

D. Hughes, Pore structure and permeability of hardened cement paste, Mag.


Cone. Res., 37, 227, 1985.

L.B.W. Jolley, Summation of Sorie_, 2nd revised ed., Dover Publications,


New York (1961).

P. Mehta, and D. Manmohan, Pore size distribution and permeability of


hardened cement paste, in 7th International Congress on the Chemistry
of Cement, 1980, Vol. 3, pp. VII 1-5.

C.E. Neuzil, C. Cooley, S.E. Silliman, J.D. Bredehoeft and P.A. Hsieh, A
transient laboratory method for determining the hydraulic properties
of 'tight' rocks II. Applieati0n, Intl. J. Rock Mech. Min. Sci. & Geomech.
Abstr. 18, 253-258 (1981)

34
B. Nyame, and J. Illston, Capillary pore structure and permeability of
hardened cement paste, in 7th International Congress on the Chemistry
of 17ement. 1980, Vol. 3, pp. VI 181-185.

J. Pommersheim and B. Scheetz, "Extension of standard methods for


measuring permeabilities of pressure pulse testing, to be submited for
publication (1989).

D. Shi, P. Brown and S. Kurtz, "A model for the distribution of pore sizes in
cement paste," 23-34, Pore Structure and Permeability of Cementitious
Materials, L.R. Roberts and J.P. Skalny, Eds., MRS (1989).

D. Shi, W. Ma and P. Brown, "Lognormal simulation of pore evolution during


cement and mortar hardening," 143-48, Scientific Ba_i_ for Nuclear
Waste Management XIII, V. Oversby and P.W. Brown, Eds., MRS
(1990a).

D. Shi, P.W. Brown and W. Ma, "Lognormal Simulation of Pore Size


Distribution in Cementitious Materials," J. Am Ceram Soc, submitted
(1990b).

D. Stauffer, Introduction tQ Percolation Theory, Ch. 2, Taylor & Francis,


London, 1985.

D. Trimmer, Design criteria for laboratory measurements of low


permeability rocks, Geophys. Res. Lett. 8 (9), 973-975 (1981).

35
Figure (_aptions

Figure 1. Schematic drawing of the pulsed permeability apparatus.

Figure 2. Exploded drawing of permeability cell.

Figure 3. Theoretical pressure-time behavior for pulsed permeability


experiment.

Figure 4. Typical raw data for pulsed permeability experiment.

Figure 5. Typical processed data for pulsed permeability experiment.

Figure 6. Cumulative pore size distribution of FA cement (w/c = 0.55, f/c =


0.536, cured for 28 days).

36
37
:! ,i

ei

38
eJnsseJd

39
0

- I0
IJJ

03
¢0
LU

0 a:
- ,_0 a.

®
X ®

0 0 0 0 0 0 0 0

40
¢D

41
42
POROSITY/PERMEABILITY RELATIONSHIPS

P.W. Brown
Department of Materials Science and Engineering
. Pennsylvania State University
Dex Shi
Civil Engineering Department
Pennsylvanic State University

J.P. Skalny
W.R. Grace

Iotroduq_ign

The structure of the porosity in concrete strongly influences its


performance. Specifically, porosity determines the rates at which
aggressive species can enter the mass and cause disruption. Rates of
intrusion are related to the permeability of the concrete. In the most
general way, permeability depends on the total porosity. More
importantly, however, permeability depends on way in which the
total porosity is distributed. Porosity, in turn, is related to the
original packing of the cement, mineral admixtures, and the
aggregate particles, to the water-to-solids ratio, to the rheology,
which is related to the degree of dispersion of the solids originally
present, and to the conditions of curing.

This contribution considers the nature of porosity in cement


and concrete, discusses its measurement and the limitations in the
interpretation of porosity data. Models developed to describe
porosity both in cement paste and concrete are reviewed. With
emphasis on permeability models developed for a variety of porous
inorganic materials, relationships between pore structure and
permeability are discussed in terms of their applicability to concrete.

The Nature of Porosity in Cement Paste. Mortar and Concrete

From a pragmatic standpoint, porosity of a material is not of


interest as an end in itself. Rather porosity is of interest because it
directly influences both mechanical and transport properties of
o cementitious materials. For example, the pore shape-dependent
relationship between porosity and strength, <_, of the form:

<_/o'o = exp-(be) [1]

*Materials Science of Concrete II (in press).

48
is well known [1], where b is a shape-dependent parameter, e the
porosity and 0o the ideal strength. This relationship is illustrated in
Figure 1 [2].

With respect to durability, the ability of concrete to resist


various forms of deterioration is often related to its impermeability.
Analysis of the sources of porosity and its connectivity can
frequently provide the means to understand the mechanisms by
which aggressive species can intrude concrete. This is because the
pore structure defines the paths along which liquid or vapor
preferentially moves. This, in turn, is of obvious importance with
respect to specific durability considerations including the transport
of freezable water and electrolytes, such as chlorides, through
concrete.

It is widely accepted that permeability is determined by


microstructure. Microstructure in this context is defined in terms of
pore and crack structures. A large number of models have been
developed, particularly for sedimentary rocks, to predict the
permeability from measurements of pore structures and cracks [3-
12]. For cementitious materials, it is well recognized that both total
porosity and its distributions determine the permeability [13-16],
and that only pores with diameters greater than a specific value
contribute significantly to permeability [13, 14, 16]. Figure 2, for
example, illustrates the dependence of permeability on the porosity
[17]. It has also been observed that the inflection point on the
cumulative pore size distribution obtained by mercury porosimetry
locates the minimum diameter of pores which form a continuous
network through hydrating cement paste [13, 14, 16].

Large variations in permeability of concrete having nominally


similar porosities are frequently observed [13-15, 18]. Verbeck even
implied that the permeability of concrete may not be a fundamental
material property [19]. Alternatively Podvaley and Prozenko [20]
claimed that the variations should in most instances be considered an
objective characteristics of the phenomenon, and that the source of
variation in permeability results from small variations in concrete
microstructure. As a consequence, it is useful to enumerate the
various types of porosity present in concrete and to establish their
relative contributions to permeability.

There are a variety of "types" of porosity in concrete. These


types may be classified in terms of their origin or in terms of their
anticipated effect on measurable parameters such as strength or
permeability. Sources of porosity in concrete include:

44
1. gel pores
2. smaller capillary pores
3. larger capillary pores
4. large voids (also included in this category may be intentionally
added voids such as by air entrainment)
5. porosity associated with paste-aggregate interracial zones
6. microcracks and discontinuities associated with dimensional
instabilities that occur during curing
7. porosity in aggregate

The diameter of a stable gel pore is assumed to be about 2 nm


[21]. The reason for the selection of this value is based on the
assumption that hydration products cannot precipitate in pores
having diameters smaller than about 2 nm. Because the gel porosity
resides in the hydration products that accumulate between the liquid
phase and the anhydrous cement grains, gel porosity has a major
effect on hydration rates but only a minor effect on transport
processes involving liquids. However, there is at present no
justification for ignoring the other types of pores listed above. Thus,
the contribution of each of the remaining types of porosity to
permeability must be considered. Unfortunately, deconvolution of
the relative contributions of each of these sources of porosity to
permeability has not been carried out. Therefore, conclusions
reached regarding concrete permeability are frequently based on
extrapolation of results obtained for cement pastes. However, it may
be reasonable to subdivide the porosity in concrete into two classes:
(1) that in the paste matrix and (2) that associated with the
aggregate and paste interface.

The principle source of the matrix porosity contributing to


permeability is that associated with residual space between cement
grains which was originally filled with water. The contribution of this
source of porosity may be amenable to assessment by investigations
of cement paste. However, the contribution of the porosity
associated with the interracial zones between paste and aggregate
and the microcracks that develop in this interfacial region, which
extend into the paste, to permeability must be assessed by
determinations carried out directly on aggregate-containing
materials.

Two types of porosity can be considered to form the network of


capillary porosity present in cement and concrete: large and small
capillary porosity. One reason for categorizing porosity in this
manner is related to the influence of chemical and mineral
admixtures on the two types. Capillary porosity is assumed to have a
major effect on transport processes but only a minor effect on

45
hydration rates. The diameters of capillary pores could in theory
range from very small values to large ones. However, it has been
assumed that the lower diameter limit of capillary porosity is 100
nm [14, 16, 22]. It is worthy of note that there is an apparent
discrepancy between the size of a gel pore (2 nm) and the lower
bound on the size of capillary pore. According to the IUPAC [23]
classification of pore sizes, the diameter of a micropore is 2 nm or
less while that of a mesopore is between approximately 2 and 50 nm.
Mesopores are of a size range for which electrostatic interactions
between the pore walls and the liquid would extend over a
significant fraction of the cross-sectional area. A consequence of this
may be that transport processes through pores having diameters in
this range are hindered by electrostatic effects. It is well known that
the mineral admixtures affect permeability; the basis for this effect
can be understood in terms of the formation of a larger amount of
porosity in the mesopore range. However, this view requires further
experimental verification for both portland-pozzolan and pc.tland-
slag systems.

Overview of Porosity Measurement and Jntcrpretation of D_ta

Three important types of techniques have been used to


experimentally measure the porosity and/or its distribution in
porous materials. These techniques are gas adsorption, mercury
intrusion, and direct observation techniques including serial
sectioning and pore casting followed by optical or SEM observation of
the sections or the casts.

Vapor or Gas Adsorption

This technique involves the adsorption of a gas (including


water vapor) on the accessible internal surfaces of a specimen.
Because a significant degree of uncertainty regarding the inter-
relationships among hydration phenomena, pore size distributions,
and properties of cement systems remain, studies using this
technique have been carried out over the past decades to elucidate
them. However, internal surface areas determined by gas adsorption
experiments are dominated bythe gel porosity, which has only a
minor effect on bulk transport.

Mercury Intrusion

A second, important technique to determine porosity is


mercury intrusion. In this technique, mercury is forced into porous
samples with increasing pressure. Because the mercury is assumed
not to wet the internal sample surfaces (the angle of contact between

46
a drop of mercury on a cement surface representative of the the
internal surface and the surface would be greater than 90°), the
intrusion pressure can be related to the diameter of the pore
intruded provided the contact angle between the mercury and the
, sample are known or can be assumed. The value assumed for the
contract angle allows the calculation of the diameter of the pore
intruded according to the Washburn equation [24]:

P = -2_cosO/r [2]

where P is the pressure needed to intrude a pore of radius r, x is the


surface tension of mercury, • is the angle of contact.

The application of mercury intrusion porosimetry to measure


the distributions of pore sizes in cements was pioneered by Winslow
and Diamond [25, 26]. Although there are certain limitations to the
use of mercury intrusion in the analysis of the pore structures of
cements, the technique is very attractive since it measures pores of
the general size range that appear to control permeability.

Data obtained by mercury intrusion tiorosimetry (MIP) must,


however, be interpreted with caution. Pores in the general size
range or 50 to 100 nm are of a size for which capillary condensation
is of importance. The occurrence of capillary condensation is based
on the assumption proposed by Zsigmondi [27] that the equilibrium
vapor pressure in a pore of a given diameter can be determined by
the radius of curvature of the meniscus of the liquid in the pore.
This leads to the familiar Kelvin equation:

ln(po/p) =-2xM/(SaRT) [3]

where Po is the saturation vapor pressure of the bulk liquid, p is the


saturation vapor pressure of the liquid in a pore of diameter a, x is
the interracial (surface) tension, M is the molecular weight, 8 is the
density, T is the absolute temperature and R is the ideal gas constant.
Rearranging the Kelvin equation gives:

p = po/(e x) [4]

where x is the right-hand term in Eq. [3], and shows that the
saturation vapor pressure in a pore is reduced with respect to that in
the bulk liquid. A consequence of this is that pores in this size range
remain filled with liquid at relative humidities below 100%.
Therefore, attempts to minimize disruption to pore structures by
drying under conditions less aggressive than 105"C, for example,
may result in incomplete desiccation.

47
Another important factor complicating interpretation of pore
size data for cements is the presence of "ink bottle" pores. Gas
adsorption experiments on cements have shown that there is
significant hysteresis between the adsortion branch and the
desorption branch of the curves [see for example 28]. Katz [29] was
probably the first to point out that sorption hysteresis occurs when
smaller capillary pores do not empty during desorption with the
effect of blocking the emptying of larger capillary pores. Generally
similar behavior is observed with mercury intrusion experiments;
some of the mercury intruded under pressure remains trapped in
cement specimens when the pressure is removed even though is it
assumed that the mercury does not wet the specimen surfaces. This
is the result of mercury being forced into small-necked pores. A
consequence of this phenomenon is that large pores with small necks
appear as many small pores. Cerbesi [3.0] reported that
approximately 75% of the pore volume of cement paste interrogated
by MIP is composed of pores uniform in cross section. The remaining
25% are "ink bottle" pores. Because bulk transport is primarily
affected by the minimum pore diameters, this phenomenon may
influence the interpretation of MIP data with respect to
permeability.

An additional factor that must be considered when pore


structures are determined by mercury intrusion porosimetry is that
closed porosity is difficult to determine unless the pore walls are
damaged by the intrusion process. It is not necessarily undesirable
that closed porosity is not intruded because it is unlikely that closed
porosity would significantly contribute to transport processes.
Alternatively, the rupture of pore walls would result in regions of
closed porosity being included in the measurement of porosity
because the membranes of hydration product that isolate these
regions from the capillary porosity network are destroyed during
mercury intrusion [see for example 31].

Another limitation to the interpretation of pore size data is


related to the uncertainty in the value of the contact angle used in
the Washburn equation. Finally, because specimens are routinely
desiccated before mercury intrusion, mesopores will appear to
contribute more to the pore size distribution of the fine fraction than
is appropriate.

Taken together, these phenomena serve to complicate the


interpretation of experiments using mercury intrusion porosimetry
which attempt to assess both the volumes and size distributions of
pores that significantly contribute to permeability.

48
Direct Observation of Porosity

There are two types of techniques used to directly observe the


porosity in cement paste, mortar, and concretes. The first of these is
the direct observation of large pores on a polished section using
optical microscopy. The observation of air void spacings is one
example of this. However, the observation of finer pores usually
involves the intrusion of a monomer and its polymerization or the
intrusion and hardening of an epoxy. Porosity can then be observed
in polished sections or in thin sections via optical microscopy [32, 33].
The observation of porosity with an optical microscope is frequently
enhanced by the use of ordinary or fluorescent dyes [34]. This
technique has also been applied to the observation of microcracks in
mortar [35]. The observation of porosity by optical techniques is
limited by the resolution of the optical microscope. Typically these
analyses are carried out at low magnification, 40-400X.

Optical microscopic observations may also be carried on


polished sections from which material is systematically removed by
cyclic polishing between observations. Analysis of the pore size data
from the serial sections can then be used to assemble a montage of
the three dimensional pore structure. A less tedious variant of the
technique of preparing serial sections is pore casting. Pore casting
has found application in the analysis of the pore systems in rocks
using both optical and electron microscopy [36, 37] but very limited
application in cements [38]. In this technique the pore structure is
filled with a plastic material and the inorganic matrix is dissolved to
reveal the three-dimensional pore network. The advantage to pore
casting is that it is possible to analyze the pore structure as a
network.

Electron microscopy has also been used to analyze the porosity


on polished sections. This is accomplished either by direct
observation of the pores or by intruding the pores with a compound
containing an element readily detected by energy dispersive x-ray
analysis, such as a chlorine-containing polymer, followed by an areal
compositional analysis for regions high in chlorine. Typically, this
can be done with a resolution of 3-5 v m.

Regardless of the specific technique selected, there are two


major limitations of direct observation of pore structures. The f'Lrst is
the limitation on resolution when optical microscopy is involved. The
second limitation is that regarding the ability to intrude epoxy or
monomer into small pores. However, regardless of these limitations,
direct observations of pore casts, polished sections, and thin sections

49
in concert with image processing have developed into a powerful
technique for pore structure analysis [39-43].

Models for Pore Size Di_ribution_ in Cements

In spite of the variety of limitations, previously discussed,


mercury intrusion has come to be regarded as a standard means by
which the pore structures of cement pastes are examined. The
pressure at which mercury will flow into a pore is related to the size
of that pore. Thus, by determining the volume of mercury intruded
as a function of intrusion pressure, plots as shown in Figure 3 [25]
are obtained. While data of this type are useful in comparing
relative pore size distributions in various cement pastes, further
analysis is required to extract descriptors of these pore size
distributions. Such descriptors can then be used in permeability
expressions.
The most common method of analyzing porosity data is to
determine an averaged pore size. This is the most straightforward
approach. However, from the standpoint of those processes that
affect cement porosity this may not be the best approach. Rather,
the nature of the pore size distributions need to be defined in terms
that can be treated quantitatively. This is most readily done by
describing pore sizes using a distribution function. The general form
of a function that may be used to describe a cumulative pore size
distribution is:

P(R >r) = V(r)dr; P(O) = 1,


fr °Q
[5]

where P(R>r) is the probability that a pore will have a radius larger
than r, i.e. the cumulative pore size distribution, V(r)dr is the volume
fraction of pore space whose radii are between r and r + dr. The
probability that a pore will have a radius larger than zero is unity.

Little appears to have been done in terms of developing


distribution functions describing the porosity of cement pastes. The
exception to this is an investigation by Diamond and Dolch [44]. In
this investigation it was reported that the distribution in pore sizes
in a hardened cement paste could be described by a log-normal
distribution function. This is a significant finding because the data
were obtained by mercury intrusion porosimetry. Diamond [45] also
measured the variations in pore size distributions with temperature
over a range from 6 to 400C. He observed pore size distributions to
be initially coarser in pastes cured at elevated temperatures but that
the differences became negligible after about 1 month of curing. In

50
accord with earlier work, Diamond also observed that the
distribution in porosity could be treated as log-normal. Recently, Shi
and coworkers [46, 47] have extended the analysis to MIP data
obtained at high intrusion pressures. Using robust statistical
techniques Shi, et al have observed that is possible to describe the
distributions in pore size in both cement paste and mortar in terms
of a linear combination of lognormal distributions. The equations
, used to describe distributions of porosity are of the form:

p(r) = Z fi p(r, _ti, ci), Z fi = 1.0 [6]

where p(r) is the probability density function, fi the weighing factor


of the itla lognormal sub-distribution, i.ti and _i the location parameter
and shape parameter in the i th lognormal sub-distribution p(r, I.ti, ai),
respectively. As hydration proceeds, the location parameters change
significantly, while the relative weights of the sub-distributions and
the shape parameter change relatively little. The significance of this
finding is that it is possible to describe the variation in porosity
distribution as a function of water-to-cement ratio and as a function
of curing time in terms of the variation of parameters having direct
statistical significance. These parameters can be incorporated, in
turn, into pore structure-permeability relationships.

Techniques Used to Measur_ P_l'mcability

A variety of techniques have been developed for permeability


measurement. A number of these are summarized in ref. [48]. The
most common technique is the measurement of the amount of water
that can be forced through a specimen subjected to a large hydraulic
head under conditions which approach steady state. Such
permeability measurements suffer from several disadvantages. If a
sample is relatively impermeable, the use of a relatively small length
is required. This raises the question of representativeness. The
solubilities of hydrated cement phases will vary in response to the
local hydrostatic pressure. Thus a pressure gradient through a
sample can lead to the redistribution of the more soluble phases,
Ca(OH) 2 and gypsum, if present. Redistribution of solids can affect
the pore structure and, thus, the permeability. A third disadvantage,
which is ubiquitous to essentially all permeability measurements, is
that specimens having low permeabilities are saturated only with
great difficulty.

A second, relatively common method for measuring


permeability is the rapid chloride permeability method [49]. In this
method, chloride ions migrate through a specimen under a potential
gradient. The advantage to the method is its relative rapidity.

51
Disadvantages are that there is resistive heating associated with the
voltage drop across the sample and that the chloride can rcac_ with
aluminam phases and influence the pore structure.

One approach to eliminating the problems inherent in the more


traditional methods of permeability measurement is to minimize the
pressure gradient across the sample. Such a method has been
developed to measure the permcabilities of rocks [50]. According to
the method, the pressure gradient across the sample is initially zero.
A zero pressure gradient is attained by using two pressurized
reservoirs. In order to make a measurement, the pressure in one of
the reservoirs is rapidly elevated or lowered and the rate of decay in
the pressure gradient across the sample is measured. The rate of
decay can then be mathematically related to the permeability [50].
While this method offers the advantages of rapidity in measurement
at relatively low pressure gradients [51], obtaining fully saturated,
low permeability samples remains a problem [52].

Models for the Pcrmeabiliti¢_ of t_¢rnent and (_oncret¢

Consideration of models developed to "describe the permeability


of porous media will include empirical models, network models,
probabilistic models and models based on percolation theory.

Empirical Models

In general, permeability is a measure of the ease with which a


fluid passes through a porous body [53]. The most common
expression used to describe the permeability, k, is Darcy's Law:

k = -p.Q/[ A 8g(dh/dz) ] [7]

where Q is the volume of fluid discharged per unit time through the
cross-sectional area A, I_ is the viscosity of the fluid, 8 is the density
of the fluid, g is the acceleration of gravity, dh/dz is the hydraulic
gradient in the direction of flow, z.

Another simple model is that of Poiseuille [4]. Poiseuille's law


states that the volume flow through a capillary tube of diameter r is

Q = -xr4/8ttdP/dl

where p. is the viscosity of the fluid and dP/dl is the pressure


gradient causing flow along a tube of length I.

52
These simple models Can be combined to develop an expression
for permeability. Assuming the porosity in a cross-section through a
material with total porosity p to result from the intersections of
pores having different diameters ri, the term A in Darcy's law can be
expressed as S/p. S is the cross-sectional pore area. In this case, the
total volume flow can be expressed in both Darcy's law and in
PoiseuiUe's law as:

Q = -S/e k/gdP/dl (Darcy's law) [9]

Q = -Z_(ri)4/SgdP/dl (PoiseuiIlie's law) [10]

(The terms g and 8 only result in different units)

Combining both equations results in:

k = e/S_;_z(ri)4/8 = e(_(ri)2(ri)2/8)/S

= e(Y_Si(ri)2/8)/S = e <r2>/8 [11]

where Si is the cross-section area of pores of radius ri, <r2> the mean
squared pore radius, or the second moment of the pore radius
distribution. This model relates permeability to porosity and the
average pore size by assuming that pores are tubes and do not
interconnect.

Another frequently encountered permeability model is the


Carmen-Kozeny model. This model is sometimes referred to as the
hydraulic radius model since it assumes pore diameters to be 4 times
the void volume of the medium divided by the pore surface area. It
relates the permeability to total porosity and the specific surfaces
area of the pores. The mathematical form of the Carman-Kozeny
model is

k = e3/[ko (Le/L)2(1 - e)2S 2] [12]

where e is the total porosity, k o is the permeability of an infinitely


dilute bed, Le is the average path length for flow, L is the path length
(Le/L is often called the tortuosity), S is the specific surface area [52].

Archie's law model [55] relates porosity and permeability using


a power law. This model is expressed as

k **eta [13]

53
being the total porosity and m a constant. In the Archie model
factors as connectivity and tortuosity, which are important to
permeability of cementitious materials, are not considered.

Each of the above models relates permeability to the porosity


in some fashion. It is of significant interest to relate the
permeabilities of cementitious materials to parameters that control
the development of microstructure. As a consequence, a variety of
investigations have been carried out to elucidate possible
relationships between pore structures and permeability in
cementitious materials. Auskern and Horn [56] compared the pore
size distributions obtained by mercury intrusion porosimetry for
pastes hydrated at water-to-cement ratios of 0.35 and 0.55. After
90 days of hydration both the distribution in pore sizes and the total
porosities were similar for the pore diameters of 90 nm and below.
This suggests that the variation in permeability with water-to-
cement ratios is dominated by the larger porosity. Nyame and
Illston [13] reported a linear relationship between the maximum
radius of a continuous pore and the permeability of a paste. Mehta
and Manmohan [14] found that the permeability of pastes could be
correlated with the volume of pores having diameters greater than
132 rim. Goto and Roy [16] observed that pores in the size range
from 75 to 750 nm have a major influence on paste permeabilities.
However, it has also been established that the techniques used to
determine the pore structures of cementitious materials may affect
the pore structures themselves. For example, Hughes [15] confirmed
the observations of Feldman [31] and others that mercury intrusion
significantly damages the pore structure of cement paste. Hughes
also reviewed the work of Marsh, which indicated the lack of any
correlation between permeability and pore size distribution, and
developed a simple tube model to predict permeability.
Unfortunately, a generic model having predictive capabilities does
not seem to have emerged from these investigations and the
capabilities of the models described above appear to be limited. At
best it appears that, while there not agreement regarding the
relationships between porosity and permeability, the weight of
evidence suggests the larger capillary porosity to primarily
contribute to the permeability in cement paste. However, care must
be taken in generalizing these observations to concrete.

P_rmeability Models

The following sections describe number of classes of models


which have not been applied to cementitious systems but that have
proven successful in predicting permeabilities of catalysts, _of

54
geological materials including rocks and soils, and of permeability
analogs in electrical systems.

Network Models

The Daxcy-Poiseuille model, the Archio model and the Carman-


Kozeny model all attempt to relate permeability to some averaged
descriptor of the porosity, or to the total porosity. They do not
consider the structures in pore network. In spite of this limitation
these models can, in some instances, qualitatively predict concrete
permeability [see for example 57, 58]. However, the predictions
based on these models are frequently considered to be inadequate
[59]. This recognition occurred in the geosciences in the 1950's and
led to the development of a "network model" to describe
permeability [60-62]. Since then, network models describing porous
systems have been developed in the areas of geophysics [63],
petroleum geology [64, 65], soils [4] and chemical engineering [65].
However, none seems to have been specifically developed in the area
of cement and concrete.

Network models axe based on the analogy of Darcy's law for


fluid flow to Ohm's law for current flow. Seeburger and Nur [59], for
example, investigated the effects of confining pressure on
permeability and bulk modulus in rocks using a network pore space
model as a tool. In their network model, the elements are capillary
tubes of given length, cross section and shape. The change of
permeability with hydrostatic confining pressure is caused by the
effect of the stress field on the shape and resulting flow
characteristics of each element. In a model developed by Dullien
[54], a network consists of many sub=networks of pores that have a
specific smallest entry pore size, Figure 4. Each sub-network is
treated as a flow channel. The volume flow in each channel is
determined by the mean squared diameter of pores in that channel.
Total volume flow is the sum of that in all flow channels. DuUien's
model [3, 54, 70] is written as foliows:

k = ¢ <D2>/96 [14]

where <D2> is the mean square pore diameter and e theporosity. The
value of the constant in Eq. [14] differs from that in the classical
. Poiseuille model, Eq. [11], by a factor of 3. Dullien interpreted this
difference as a tortuosity factor. It is doubtful, however, whether
the tortuosity factor is a constant. A tortuosity factor is more likely a
variable depending on microstructure. The mean square pore
diameter in the model is determined from a bivariate pore diameter
distribution proposed by Dullien [71], as will be discussed as follows.

55
Mercury intrusion porosimetry can only determine pore entry
sizes. Quantitative image analysis can determine the "true" pore
sizes, although only in two dimensions. If the size distribution
curves determined by these methods overlap over the entire range
of pore sizes, a bivariate pore size distribution can be expressed as
p(D, De) where D is true pore diameter and D e is pore entry diameter,
Figure 5. Dullien developed a simple, graphical procedure to
estimate the volume of pores having the true diameter D to D + dD
and the entry diameter De to De + dDe. A bivariate distribution was
found to be suitable to account for both distributions [3, 57, 71].
From this bivariate distribution, Dullien claimed that the relation
between pore entry diameter and pore diameter can be determined.
It must be noted, however, that the bivariate distribution cannot be
established unless two distributions overlap over the whole range of
pore sizes. In less porous materials, such as cement and concrete,
usually there are many fine pores, which are not amenable to
observation by image analysis. In addition there are very large
pores (voids) which can be difficult to measure by mercury
porosimetry. Thus, overlap in the whole range of porosity may not
be possible. It is also worth noting that the pore size distribution
determined by image analysis represents the size distribution of the
intersections of pores by a plane. Three-dimensional, or the true
volume distribution of pores must be estimated using stereological
procedures [72-77].

Probabilistic Models

Probabilistic models were first developed by Childs and Collis-


George [67]. Subsequently, models have been developed by Marshall
[68] and others [4, 69]. The basis for a probabilistic model is that a
porous solid can be divided into two parts by a plane normal to the
flow direction and that the two cross-sections can be rejoined. If
pores are distributed at random and the term f(D)dD is the
probability that a pore has diameter within the range D to (D +dD),
then the probability that pores with diameter Di to Di + dDi on cross-
section i is connected to pores of size Dj to Dj + dDj on cross-section j
is f(Di)f(Dj)dDidDj. If the interconnection between pores of the two
cross-sections is assumed to be completely random, the permeability
coefficient can be expressed as:

k **e2j'j"D2f(Di)f(Dj)dDidDj [15]

where e is the porosity and D the smaller of Di and Dj. The


proportionality factor is 1/32 if circular capillary tubes are assumed
[4]. Juang and Holtz [4] noticed that in reality, the connection

56
between pores of two cross-sections would not be completely
random. They added a factor called the "connection function" g(y, Di,
D j), where y is the sample length, to control the connectivity and
realized that it accounts for tortuosity of flow. When y_ 0, it
, becomes k = e2<D2>/32, a form of the classical Poiseuille capillary
model [4]. As Gueguen and Dienes pointed out [6], the tortuosity
concept is related to percolation concept, as will be discussed

Models Based on Percolation Theory

In 1957, Broadbent and Hammersley introduced the term


"percolation theory" to describe critical phenomenon, or phase
transition behavior [78]. The basis for percolation theory can be
considered in terms of the site occupancy of a lattice. A primary
assumption is that every site is either "occupied" or "empty" based
on an entirely random process which is independent of the
occupancy of neighboring sites. The probability that a site is
occupied is p and (I - p) is the probability that a site is empty.
Occupied sites are either isolated from each other or adjacent to
neighboring sites to form clusters. If the site occupancy is near zero,
most occupied sites will be isolated. If, °n the other hand, the site
occupancy is close to unity, then almost all occupied sites will be
connected to form a large cluster, which extends from one side of the
lattice to the other. This large cluster represents an "infinite-path."
It percolates through the lattice in the similar way that fluid
percolates through porous materials along the network of connected
pores. Increasing the site occupancy from zero to unity results in a
critical point (percolation threshold), Pc, above which an "infinite-
path" can form. The occurrence of an infinite path is illustrated in
Figure 6 for a simple model of fluid flow through a porous medium.

The percolation described is called "site percolation" [79-83].


There is a counterpart called "bond percolation" [79-83]. In this
instance, each site on the lattice is occupied, and interconnections
between neighboring sites are regarded as bonds. In this instance, p
is the probability that a bond is "open" and (l-p) is the probability
that a bond is "closed." A cluster is a group of neighboring sites
connected by open bonds. The percolation threshold, Pc, is the
fraction of open bonds, below which an 'infinite-path' cannot form.
Fluid flowing through porous materials can be modeled in terms of
its passage through a network of interconnected pores.

Intermediate between site percolation and bond percolation


there is a type of percolation called "site-bond percolation" [79, 80,
84]. The lattice sites are no longer all occupied as in bond
percolation. In site-bond percolation, p is the probability that there

57
is a bond between neighboring sites. A cluster is then a group of
neighboring occupied sites connected by bonds. The bond
percolation threshold decreases from unity, when the portion of
occupied sites equals the site percolation threshold, to the normal
bond percolation threshold, when the fraction of occupied sites is
unity. This type of percolation appears to most closely represent the
nature of fluid flow through porous materials. The fraction of
occupied sites is equivalent to porosity, the bonds represent the
channels connecting pores, and a cluster represents "infinite-path"
pores. The channels may consist of pore entries and cracks.

In the percolation theory, P is defined as the strength of


infinite network. The strength, P, is the percolation probability that
a site belongs to the infinite-path. It is actually the fraction of total
number of lattice sites that are on an infinite network [79]. A power
law expression may be written:

P ** (p - pc) t [16]

if p approaches Pc from above where p is the concentration of


occupied site or bonds, and Pc is the critical concentration. Transport
properties are proportional to (p - pc) cx. ct is the transport exponent.
However, t usually differs from a [85] because P counts for both
backbone and dead ends. Dead ends contribute to the mass of the
infinite network but not to transport properties. However, dead ends
entrap fluid.

From the standpoint of flow in porous materials, an infinite


network consist of two parts: backbones and dead ends. Fluid flows
through a porous materials along the backbones and are entrapped
in dead ends. Permeability measurement usually only deals with the
flow-through fluid; however, entrapped fluid can significantly affect
properties, such as freeze-thaw resistance of concrete. Therefore, it
is necessary to consider phenomena associated with flow in both
backbones and dead ends. Computer simulation can be used to
differentiate dead ends from backbones and to calculate fractions of
backbone and dead ends with respect to the infinite network and
numerical studies have been made to quantify backbones [86].

Several of the models previously described above can be


considered in terms of percolation concepts. Both percolation theory
and Archie's law use a power law to describe the critical
phenomenon. Thus, Archie's law may be regarded as a special case
in which the percolation threshold is zero [87]. For dense yet
permeable materials, the percolation threshold can approach-a very
low value. For example, rock salt has porosity of only 0.6% but a

58
permeability of 7.3 x I0-6 Darcy [88]). Dullien observed the
permeability of sandstone to mercury to depend on the degree of
pore prefilling by mercury. That is pores of a given size that were
previously filled with mercury contributed to permeability while
. those that were not filled did not contribute to bulk flow. Dullien
concluded that at a low mercury saturation there must be continuous
flow channels in the medium. At low saturation, only the largest
entry pores are penetrated by mercury. In other words, these flow
channels consist only of pores with large entries that can be intruded
by the mercury. However, if the saturation is reduced to a very low
level, no permeability is recorded [3]. This low saturation must be
related to percolation threshold. Thus, Dullien's model as well as
Juang's statistical model can be regarded as transitional models
between non-percolation-based models and percolation-based
models.

Gueguen and Dienes [6, 89, 90] used statistics in combination


with percolation theory to correlate pore and crack structures with
permeability. Their models assume narrow size distributions and
can be expressed as:

k = fE<r>2/32, for pore structure and


[17]
k = 2fe<w>2/15, for crack structure,

f is the fraction of connected pores and is proportional to (p - pc) a,


<r> is the mean pore radius and <w> is one half the mean crack
aperture, and ¢ is the porosity. Unfortunately, the assumption of
narrow size distribution is not valid for cementitious materials.
However, the model explicitly includes f, the fraction of connected
pores and thereby provides a means to quantify the tortuosity factor
[6]. Additionally, their model provides insight to crack-permeability
relationships; in cementitious materials pores and cracks are not
distinct and the effect of cracks on permeability must be considered.

Katz and Thompson [5, 91, 92] developed a percolation model


using an approach based on the work carried out by Ambegaokar,
Halperin and Langer (AHL) [93] on electron hopping in amorphous
semiconductors. They showed that transport in a random system
with a broad distribution of conductances is dominated by those
: conductances with magnitudes greater than some characteristic value
go. gc is the largest conductance such that the set of conductances
; {g:g > go) forms an infinite network. Transport in such a system
reduces to a percolation problem with threshold ge- Kirk-patrick [82]
and Shante [94] extended these ideas and assigned the value ge to all
local conductances with values g >-gc and set all conductances with

59
values g < gc to zero. They arrived at a trial solution for the sample
conductance of the form:

g = agc[p(gc)
-Pc]a, [18]

where P(gc) denotes the probability that a given conductanceis


greaterthan or equal to gc, and a is a constant.Pc is the critical
concentration of conductance, below which no infinite cluster can
form. The transport exponent, cc is approximately 1.9 [95, 96]. This
model is similar to a site-bond percolation model. Katz and
Thompson [5, 92] applied these concepts in the development of a
model describing the permeabilities of porous rock. They defined
threshold conductance value gc as function of the characteristic
length 1c" gc oo lc. Assuming cylindrical pores, the characteristic
length is the pore diameter. Then the hydraulic conductance of the
sample is the function of the length parameter 1:

g(1) - Egc(1)[p(1) - pc] ct [19]

where e is the porosity. As 1 is lowered past the threshold 1c the


function g(1) also decreases but the power-law term [p(1) pc] 0_
increases as more and more pores are included in the infinite
network. Mathematically, this means that there is a maximum value
for 1 that places a lower limit on g(l). Physically, because only
conductances with characteristic lengths I >_.Ic are included in the
calculation, the sample conductance, g(lmax), must be always less
than or equal to real sample conductance. In other words, g(lmax)
approaches the true conductance as closely as possible. After
maximizing, the model is expressed as follows:

k = (1/226)1c a/ao, [20]

where 1c is the threshold characteristic length, which is inflection


point on the cumulative pore size distribution determined by
mercury porosimetry. _ is the electric conductivity of the rock, and
ao is the conductivity of the pore solution. They also proposed that
electrical conductivity can be determined directly from mercury
injection [91, 92]. The result is a pore structure-permeability model
which can predict permeability simply from mercury intrusion data
as follows:

k ** (lmax)2(lmax/lc)ef(lmax), [21]

where f(lmax) is the volume fraction of pores intruded by mercury.


lmax is the maxima of 1. The advantage to this percolation-based
model is that the use of percolation factors that are difficult to

60
determine, such as those used in Gueguea's model (f, p, and pc), is
avoided. The disadvantage is that the model is based on the
maximizing a lower bound [97]. As a consequence, the calculated
permeability is typically less than the measured value. Another
disadvantage to this model is that mercury porosimetry may not be
able to describe crack structures correctly, although crack structure
may be as important as pore structure in terms of their effects on
' permeability. Garboczi has recently applied Katz and Thompson
model to cement and obtained encouraging results [968], however
permeability and pore size distribution data have not yet been
obtained on the same samples.

61
tmmaz.
A variety of models have been developed to predict
permeability from pore structure. A common feature of these
models is that a power law may exist between permeability and
some characteristic pore dimension. The tortuosity factor has been
included explicitly in J'uang's model, Dullien's model (though it is
assumed as a constant) and all percolation-based models (in the form
of percolation factors).

Each model has its rational base. Statistical models consider all
pores, whereas the Katz-Thompson model only considers large pores.
Dullien's work provides the the basis for relating entry pore volume
and true pore volume. Gueguen and Dienes have considered crack-
permeability relationship and explicitly used f, the fraction of
connected pores in their-models. Alternatively, Katz and Thompson
avoid calculating f, which is difficult. Rather, they predict
permeability from mercury porosimetry data. Even the traditional
Carmen-Kozeny model has its advantages. This pore structure-
permeability model requires as input is a pore size distribution.
Although a pore size distribution is more" easily determined than
permeability per se, it usually assumes shape for pores. The
Carman-Kozeny model uses the specific surface area of pores as
input. This does not require any shape assumption and can,
therefore, avoid a source of error.

The determination of permeability is of obvious importance in


a variety of disciplines. Investigators in these various disciplines
have developed both methodologies for the determination of
permeability and models to describe their results relatively
independently. It has been the objective of this review to briefly
describe measurement methods and models, along with their
advantages and limitations, that may have relevance to cementitious
systems.

62
References

1. C.J. Periera, R.W. Rice and J.P. Skalny, "Pore Structure and Its
Relationship to Properties," in Pore Structure and Permeability
of Cementitious Materials, L.R. Roberts and J.P. Skalny, Eds., MRS
(1989).

2. G.J. Verbeck and R.A. Helmuth,

3. F.A.L. Dullien, "New network permeability model of porous


media," AIChE J., 21, 299 (1975).

4. C.H. Juang and R.D. Holtz, "A probabilistic permeability model


and the pore size distribution function," Int. J. Num. & Anly.
Methods in Geomech., 10, 543-53 (1986).

5. A.J. Katz and A.H. Thompson, "Quantitative predication of


permeability in porous rock," Phys. Rev. B, 34, 8179 (1986).

6. Y. Gueguen and J. Dienes, "Transport properties of rocks from


statistics and percolation," Math. Geology, 21, 1 (1989).

7. J.R. Banavar and D.L. Johnson, "Characteristic pore sizes and


transport in porous media," Phys. Rev. B, 35, 7283 (1987).

8. B.I. Halperin, et al., "Difference between lattice and continuum


percolation transport exponents," Phys. Rev. Lett., _4, 2391
(1985).

9. B.F. Swanson, "A simple correlation between permeabilities and


mercury capillary pressures," J. Pet. Tech., 33. 2498, (1981).

10. J. Koplik, et al., "Conductivity and permeability from


microgeometry," J Appl. Phys., 56, 3127 (1984).

1 1. P.Z. Wong, et al., "Conductivity and permeability of rocks," Phys.


Rev. B, 30, 6606 (1984).

12. J.G. Berryman and S.C. Blair, "Use of digital analysis to estimate
fluid permeability of porous materials: Application of two-point
correlation functions," J. Appl. Phys., 60, 1930, (1986).

13. B.K. Nyame, and J.M. Illston, "Capillary Pore Structure and
• Permeability of Hardened Cement Paste," 7th Intl. Cong. Chem.
Cem., 3. VI-181-85 (1980).

63
14. P.K. Mehta and D. Manmohan, "Pore Size Distribution and
Permeability of Hardened Cement Pastes," 7th Intl. Cong. Chem.
Cem., L VII 1-5 (1980).

15. D.C. Hughes, "Pore Structure and Permeability of Hardened


Cement Paste," Mag. Concr. Res. 37(133), 227-33 (1985).

16. S. Goto and D.M. Roy, "The Effect of W/C on Curing Temperature
on the Permeability of Hardened Cement Paste," Cem. Concr. Res.
11(4), 575-79 (1981).

17. A. Neville, Properties of Concrete, Pitman, London (1981).

18. N. Banthia, "Water permeability of cement paste," Cem. Contr.


Res. 19, 727, 1989.

19. G. Verbeck, "Pore Structure," in ASTM STP 169B, Am. Soc.


Testing & Mat., Philadelphia, (1978)

20. A.M. Podvaley and A.M. Prozenko, "An investigation of the


permeability of porous bodies on a mathematical model,"

21. Powers, T.C., J. Res. PCA 3(1), 47-55 (1961)

22. K. Wesche, V. Herman, J.W. Weber, "On Some Concrete Properties


as a Function of the Pores in Hardened Cement Paste," pore
Structure and Properties of Materials, Part III, D-177-187,
Academia, Prague (1973).

23. IUPAC Manual of Symbols and Terminology, Appendix 2, Part 1,


"Colloid and Surface Chemistry," Pure Appl. Chem 31, 578
(1972).

24. E.W. Washburn, Proe. Nat. Acad. Sci. 7. 115 (1921).

25. D. Winslow, and S. Diamond "A Mercury Porosimetry Study of


the Evolution of Porosity in Portland Cement," J. Matls. 5(3),
564-85 (1970).

26. D. Winslow, and C.W. Lovell, "Measurements of Pore Size


Distribution in Cements, Aggregates and Soils," Powder Tech 29.
151-65 (1981).

27. R. Zsigmondy, Z. anorg. Chem. 71. 356 (1911).

64
28. S. Brunauer, J. Skalny, and I. Odler, "Complete Pore Structure
Analysis, Pore Structure and Properties of Mat¢dals, Part I, C-3-
26, Academia, Prague (1973).

29. S.M.Katz,J.Phys.Coil.Chem. 53_ I166(1949).

30. O.Z. Cerbesi, "Pore Structure of Air-Entrained Hardened Cement


, Paste," Cem. Concr. Res. 11(3), 257-65 (1981).

31. R.F. Feldman, "Pore Structure Damage in Blended Cements


Caused by Mercury Intrusion," J. Am. Ceram. Soc. 67, 30-33
(1984).

32. G.V. Chilingarian, C.Y. Zhang, C.Y., M.Y. A1-Bassam, and T.F. Yen,
"Notes on Carbonate Reservoir Rocks, No. 4: Determination of
Permeability of Carbonate Rocks from Thin-Section
Analysis,"Energy Sources 8(4), 369-80 (1986).

33. C. Straley and M.M. Minnis "Epoxy Rock Replicas for


Microtoming," J. Sed. Petro. 53(1-2), 667-669 (1983).

34. R.M. Gies, "An Improved Method for Viewing Micropore Systems
in Rocks with the Polarizing Microscope," JSPE (1984).

35. L. Knab, H. Walker, J. Clifton, and E. Fuller, "Fluorescent Thin


Sections of the Observed Fracture Zone in Mortar," Cem. Concr.
Res. 14, 339-44 (1984).

36. E.D. Pittman and R.W. Duschatko, "Use of Pore Casts and
Scanning Electron Microscope to Study Pore Geometry,* J. SOd.
Petro. 40(4), 1153-57 (1970).

37. N.C. Wardlaw, "Pore Geometry of Carbonate Rocks as Revealed


by Pore Casts and Capillary Pressure," Bull. Am. Assoc. Pet. Geol.
60(2), 245-57 (1976).

38. L. Parrott, R. Patel, D. Killoh, and H. Jennings, "Effect of Age on


Diffusion in Hydrated Alite Cement," J. Am. Ceram. Soc. 67(4).
233-37 (1984).

39. C. Lin, and M.H. Cohen, "Quantiative Methods for Microgeometric


' Modeling," J. Appl. Phys. 53(6), 4152-4165 (1982).

; 40. C. Lin, "Shape and Texture from Serial Contours," Mathematical


Geology 15(5), 617-33 (1983).

65
41. C. Lin and J. Hamasaki, "Pore Geometry: A New System for
Quantitative Analysis and 3-D Display," J. Sed. Petro., 53. 670-72
(1983).

42. C. Lin, G. Pirie and D.A. Trimmer, "Low Permeability Rocks:


Laboratory Measurements and Three-Dimensional
Microstructural Analysis," J. Geophys. Res. 9!(B2), 2173-81
(1986).

43. E.T. Czarnecka and J.E. Gillott, "A Modified Fourier Method of
Shape and Surface Texture Analysis of Planar Sections of
Particles," JTEVA 5_(4), 292-98 (1977).

44. S. Diamond and W. Dolch, "Generalized Log-Normal Distribution


of Pore Sizes in Hydrated Cement Paste," J. Coll. Interface Sci.
38(1), 234-44 (1972).

45. S. Diamond, "Pore Structure of Hardened Cement Paste as


Influenced by Hydration Temperature," Pore Structure and
Properties of Materials, Part I, B-73-88, Academia, Prague
(1973).

46. D. Shi, et al, "A model for the distribution of pore sizes in cement
paste," 23-32 in Pore $tr.¢¢ture and Permeability of
(_¢mentiti0.u_ Material_, MRS (1988)

47. Shi, D. et al., "Lognormal simulation of pore evolution during


cement hardening," in Scientific Basis for Nuclear Wast¢
Management XlII, V. Oversby and P.W. Brown, Eds. (1990).

48. p.ermeal_ility of Concrete, D. Whiting and A. Walitt, Eds., ACI SP-


108, American Concrete Institute, Detroit (1988).

49. D. Whiting, "Rapid Determination of the Chloride Permeability of


Concrete," FHWA Rept. RD-81/l19, Federal Highway
Administration, Washington DC (1981).

50. W.F. Brace, J.B. Walsh, and W.J. Frangos, "Permeability of Granite
Under High Pressure," J. Geophys. Res. 73, 2225-36 (1968).

51. R.D. Hooton and L.D. Wakely, "Influence of Test Conditions on


Water Permeability of Concrete in a Triaxial Cell," in pore
Structure and Permeability 0f Cementitigu8 Materials. 157-64
L.R. Roberts and J.P. Skalny, Eds., MRS (1989).

52. B. Scheetz, private communication (1990).

6(3
53. S. Davis, "Porosity and Permeability of Natural Materials," Ch. 2
in Flow Through Porous Media, R.J.M. De Weist, Ed., Academic
Press, NY (19(_9).

54. F.A.L. Dullien "Single Phase Flow Through Porous Media and
Pore Structure," Chem. Eng. J., 10. 1-34 (1975).

55. G.E. Archie, "The electrical resistivity log as an aid in


determining some reservoir characteristics," AIME Trans., 146,
54 (1942).

56. A. Auskern, and W. Horn, "Capillary Porosity in Hardened


Cement Paste," JTEVA 1(1), 74-79 (1973).

57. J. Vuorinin, "Applications of Diffusion Theory to Permeability


Tests on Concrete Part I: Depth of Water Penetration into
Concrete and Coefficient of Permeability," Mag. Contr. Res.
37(132), 145-52 (1985).

58. J. Vuorinin, "Applications of Diffusion Theory to Permeability


Tests on Concrete Part II: Pressure-Saturation Test on Concrete
and Coefficient of Permeability," Mag. Contr. Res. 37(132), 153-
61 (1985).

59. D.A. Seeburger and J. Nur, "A Pore Space Model for Rbek
Permeability and Bulk Modulus," Geophys. Res. 89(B1), 527-36
(1981).

60. I. Fatt, "The Network Model of Porous Media I. Capillary


Pressure Characteristics," Pet. Trans AIME 207, 144-59 (1956).

61. I. Fatt, "The Network of Porous Media II. Dynamic Properties of


a Single Size Tube Network," Pet. Trans AIME 207, 160-63
(1956).

62. I. Fatt, "The Network of Porous Media III. Dynamic Properties of


Networks with Tube Radius Distribution," Pet. Trans. AIME 207.
164-81 (1956).

63. M. Rink and J.R. Schopper, "Computations of Network Models of


Porous Media," Geophys. Prospect. 15. 262 (1967).

64. R.W. Ostensten, "Microcrack Permeability in Tight Gas


Sandstone," SPEJ, 919-27 (1983).

67
65. J.H. Thomeer, "Air Permeability as a Function of the Three Pore-
Network Parameters," JPT, 809-14 (1983).

66. A.C. Payatakes, C. Tein, and R.M. Turian, "A New Model for
Granular Porous Media: Part 1. Model Formulation," AIChE J.
19(1) 58-67 (1973).

67. E.C. Childs and N. Collis-George, "The permeability of porous


materials," Proc. Royal Soc., London, 201A, 392 (1950).

68. T.J. Marshall, "A relation between permeability and size


distribution of pores," J. Soil Sci., 9, 1 (1958).

69. R.J. Millington and J.P. Quirk, "Permeability of porous media,"


Nature, 183, 387 (1959).

70. F.A.L. Dullien, "Prediction of tortuosity factors from pore


structure data," AIChE. J. 21, 820 (1975).

71. F.A.L. Dullien, "Bivariate pore-size distributions of some


sandstones," J. Colloid and Interface Sci., 52, 129 (1975).

72. D. Shi, "Discussion on estimation of 3-D properties from 2-D


measurements without shape assumption", in Advances in
C_men_ Manufacture and Use, E. Oartner, Ed., Engineering
Foundation, New York (1989).

73. D. Shi, Ph.D. Thesis, Purdue University (1987).

74. D. Shi, "A probabilistic approach to evaluate distribution of


particles/pores without shape assumption," J. Microscopy,
submitted.

75. E.E. Underwood, Quantitative Stereology, 1-48, Addison-Wesley,


Reading, Massachusetts (1970).

76. DeHoff, R.T., The statistical background of quantitative


metallography, in Quantitativ_ Microscopy. R.T. DeHoff and F.N.
Rhines, Eds., McGraw-Hill, New York (1968).

77. E.E. Underwood, "Surface area and length in volume," in


Quantitative Microscopy. Ed. R.T. Dehoff and F.N. Rhines, Eds.,
McGraw-Hill, New York (1968).

78. J.M. Ziman, Model_ of Disorder, 370-386, Cambridge University


Press, Cambridge (1979).

68
79. D. Stauffer, Introduction to. Percolation Theory, Ch. 2, Taylor &
Francis, London (1985).

80. D. Stauffer, Scaling Theory of Percolation Clusters, Ch. 1, North-


Holland, Amsterdam (1974).

, 81. R. Zallen, The Physics of Amorphous Solid_, Ch.4, John Wiley,


New York (1983).

82. S. Kirkpatrick, in Ill-Condensed Matter, R. Balian, et al. Eds., Ch.


5, North-Hollan, New York (1979).

83. R. Zallen, in Fluctuation Phenomena, Ed. Montroll, E.W. and


Lebowitz, J.L., Ch. 3, Horth-Holland, Amsterdam (1987).

84. D. Stauffer, "Percolation and cluster size distribution," in O n


Growth and Form, H.E. Stanley and N. Ostrowsky, Eds., Martinus
Nijhoff, Boston, (1986).

85. D. Stauffer, Introduction to Percolation Theory_, Ch. 5, Taylor &


Francis, London (1985).

86. S.S.Manna, "Structure of backbone perimeters of percolation


clusters," J. Phys. A: Math. Gen. 22, 433 (1989).

87. I. Balberg, "Exclude-volume explanation of Archie's law," Phys.


Rev. B, _3, 3618 (1986).

88. S.N. Davis, in Flow through Porous Media, R.J.M. De Wiest, Ed.,
v

Ch.2, Academic Press, New York (1969).

89. J.K. Dienes, in Is_o¢_ in Rock Mechanics, R.E. Goodman and and
F.E. Heuze, Eds., Ch. 9, Ame. Inst. Mining, Metall. & Pet. Engr.,
New York (1982).

90. Y. Gueguen, et al., "Models and time constants for permeability


evolution," Geophys. Res. Lett., 13, 460 (1986).

91. A.J. Katz and A.H. Thompson, "Prediction of rock electrical


conductivity from mercury injection measurements," J. Geophys
Res., 92, 599 (1987).

92. A.H. Thompson et al., "The microgeometry and transport


properties of sedimentary rock," Adv. in Phys., 36, 625, (1987_).

69
93. Ambegaokar, V. et al., "Hopping conductivity in disordered
systems," Phys. Rev. B, 4_, 2612, (1971).

94. V.K.S Shant, "Hopping conduction in quasi-one-dimentional


disordered compounds," Phys. Rev. B, 2597 (1977).

95. R. Fisch and A.B. Harris, "Critical behavior of random resistor


networks near the percolation threshold," Phys. Rev. B, 1_, 416
(1978).

96. D. Berman, et al., "Conductances of filled two-dimensional


networks," Phys. Rev. B, 33, 4301, (1986).

97. Le Doussal, "Permeability versus conductivity for porous media


with wide distribution of pore sizes," P., Phys. Rev. B, 39, 4816
(1989).

98. E.J. Garboczi and D.P. Bentz, "Analytical and Numerical Models of
Transport in Porous Cementitious Materials," 675-81 in
Scientific Ba_i_ for Nuclear Waste Management XIII, V. Oversby
and P.W. Brown, Eds. (1990).

LIST OF FIGURE CAPTIONS

Figure 1. The relationship between capillary porosity and compressive


strength [2].

Figure 2. The dependence of permeability on the capillary porosity [17].

Figure 3. Cumulative pore volumes intruded depending on age for a Type


I portland cement paste hydrated at 24C at a water-to-cement ratio of 0.4
[25].

Figure 4. An illustration of one element in a pore network in which pore


sizes may vary [54].

Figure 5. A comparison of the distribution in the diameters of pore entries


and pore sizes [73].

Figure 6. An illustration showing the formation of an infinite path. This


illustration represents, in two-dimensions, flow paths across a bulk sample
in which a pore structure is defined by the application of percolation
theory. The hydraulic gradient is from left to right. _"

70
28

24 1_

0 I I
1O0 80 60 40 20 0

Capillary porosity (%)

71
100 f
90
"to 80 ,, ,
!

7o
" O0 "

50 /
40 , /

•_ 20 ,_,

0
U
0 10 20 30 40
Capillary Porosity-per cent.

72
I I i
I I 1
I I I
i I I
I I I
i I
I ' I i

T ,---I , 'i w I
i I ' I I ' I
I I I
I I I

I CAPILLARY ELEMENT ,I II

OF NETWORK

73
1400 " G O PORE ENTRY DIAMETER
_ [ I! DISTRIBUTION (De)
m / J_ &. -=. COMPLETE PORE SIZE

1200r. _ DISTRIBUTION (D'

" i '?f'tt "

0 20 40 60 80 100 120
PORE DIAMETER, D OR De (IJn'l)

74
_ ..... _ °_oO o_o o_Od oo ..o _.° o • • ., _o o_o _ .... o._. • _o° • •

, o o oQo o °, • .° • _ °° • _° • • o. ....

_ _!i!
___° iJl_i!_!_!_i_
i_ii_!ili_iilli!_i_iii_iiiiii_i_iiii_
ooo.

_!_.°.._!_._.i_i_._i_._ ._i._!_'_.._._.!i_'i_ i_"_ii_i_ _

• : _i _i_!_ _ _i'°_i_'" _ .! _i!_i°_._.i_.._ o_!.:_i_.i_i_i _

_°_
liiii,i i iiiii iiii!i t!ii !i
°o,_ .oo° • Oo. • °o _'._ _'_

75
RELATIONSHIPS BETWEEN PERMEABILITY, POROSITY,
DIFFUSION AND MICROSTRUCTURE OF CEMENT PASTES,
MORTAR, AND CONCRETE AT DIFFERENT TEMPERATURES

DELLA M. ROY, MaterialsResearchLaboratory,The Pennsylvania State University,


University
Park,
Pennsylvania 16802.

ABSTRACT

Permeabilitiestowater anddiffusionofionic species incementitiousgrouts,pastes


and
mortarsareimportant keystoconcrete durability.Investigationshavebeenmadeofnumerous
materialscontaining Donland andblendedcements, andthose withfine-grainedfdler,
atroom
temperatureandafterprolongedcuring at severalelevatedtemperatures up to 90"C. These
constitutepartof studiesof fundamentalmaterialrelationshipsperformedin orderto address
thequestionof long-termdurability. In general,thepermeabilit_esof thematerialshavebeen
found to be low [many <10"8 Darcy (10"13 m.s'l)l after curing for 28 days Orlongerat
temperatures upto 60"C. Theresultsobtainedat90"Caresomewhatmorecomplex.In some
setsof studiesof blendedcementpasteswith w/c varying from 0.30 to 0.60 andcured at
temperatures upto 90"Cthemoreopen-poresu'ucture (at theelevatedtemperatureandhigher
w/c) as evident from SEM microstructuralstudiesas well as mercury porosimetryare
generallycorrelatedalsowith a higherpermeabilityto liquid. The degreeof bondingand
permeability evident inpaste ormortar/rock interfacial studies
present somewhatmore
conflictingresults.Thebondstrength (tensile
mode)hasbeenshowntobeimproved insome
materials withincreased temperature. The results ofpermeability studiesofpaste/rock
couplesshowexamples withsimilarlowpermeabilities, andsomewithincreased permeability
withtemperature.

Ionic
diffusion
studies
alsobring
important
bearingtounderstanding
theeffect
ofpore
su'ucture.
The best
interrelationships
between
chloride
diffusion
andporestructure
appear
to
relate
diffusion
ratetomedianporesize.Similar
results
werefoundwiththeAASHTO
"chloride
permeability"
test.

INTRODUCTION

The_ ofhaxdened
cementpaste
tochemical
attack
andphysical
degradzdion
is
theresult_l_,ed composition and microstructuralfactors with environmentalexposure
conditions.t'-t,:j Farm's conwalling the initial,,n_'ostructure developmentaretherefore
stronglyrelatedto eventtutlconcretedurability.t";J
Tim tesismn_ of cementirlonsmaterials to chemical attack
andphysicaldegradation
is
onlyindite_y relatedtothemechanical su'ength of thematerial,
i.e.,
strongmaterials
do not
alwtysresistattackanddisintegrativeforcesinanyenvironment inwhichtheyareplaced.
Nevertheless_itisu'uethathigherstrength
concretes generally
havelowpet,
meabillties,
and
thereby
will provide bener resistance
tochemical attackandphysicaldegradation
than
Iow-su'engthmaterials.Thewater content(andw/cratio) ofacementpaste isprobably
the
mostimportant single factorindeterminingthesubsequent porosity
andpermeability
ofthe
" hardenedpaste,anddecreasing thewatercontentboth decreases permeabilityandporosity,
andusuallyincreases the strength.

Theporosity andporesnmcnn'e ofcementpastes exert majorconn'ol ontheingress of


potentially
deleterious substances. The porestructureandmicrostrucmm constitutemajor
rate-controilingfactorsfor potentially harmfultransportprocesses, whether the s_s being
tran_ iswater (togiverise toaficali-siiica
expansion),sulfate ions(causing sulfate
attack),
chloride ions(todeR._ivate theprotectivelayer on reinforcingsteel),
acids,or
oxygenorcarbon dioxide ttZl.

76
The factors
controlling
microstructure
andporestructure
havenotbeenfully
discerned
frompastknowledge.
E.g..certain
supplementary
cementingmaterials
orpozzolanas"
promotedevelopment
ofa fine
porestructure
_ resultinlowerp_,,e,
ability
(anddiffusivity
ofpotentially
deleterious
species)
farbeyondthat
whichwouldbepredicted
on thebasis
ofthe
water/cementirious
solids
ratioalone.
[12"
14]

Thediffusion ofionsandtherate ofpermeation offluids acrossaporous ccmendtious


matrix are, _uitively. relatedtothevolumeofpores andthesize andinterconnectivity of
these pores.tOJThe basisfor such a statementlies in the fact thatif these two factors were not
related to the pores in a porousmatrix but to the solidpart,the observed rates of diffusionof
_;10"i2t_/._c a!_ temperaturein porous cen'_ntitious.materialswould instead resemble
thatof aton'ncdiffuslon m polycrystallineceramicsof -I0 "'_m'/sec or lower. The
permeabi_ty ofwater through watersaturated cementitious materials hasbeenshowntobe
relatedtotheporosity andporesizedistribution. [1.21Inthecaseofionic diffusion in
water-sann'ated porous cementitiousmaterials, thediffusion coefficientsformoderately-
sized,singly-charged cationsgenerallyrangebetween I0_ IIandI0"_3 m2/sec [3"61while the
diffusion coefficients forthese ionsinpurewater is-I0""m'_/sec, tlj

Theobvious questionthusarises astowhy thediffusion isretardedthrough thematrix.


One ofthefactorscontributingtothedeceaseisthesmaller areathroughwhichtheions may
migr_ in a porousmauix viz. the areaof the pores. This areafactormay explain a two- or
thn_-fold
clane.aseinthediffusion coefficientofanionacross aporous rnau'ix
ascompared to
that
inthepu_ aqueous phase. Itwas pointed outearlierthat theobservc_lreductions are
generally
muchl_g_p',d_pn two-orthree-fold, inthevicinity ofI0_toI0'* times. Considering
recent thep_bable re&sonforsucha large
evMuadonsr'.J,'-'J retardation
arisesfromthe
consu'iction
andtormosityoftheporepaththrough whichanionmigrates ina water-saturated
porousroan'ix;
Theion-wall interaction leads toa substantialdecreaseindiffusion. Recent
papershaveelucidatedthis
point.[I0.III

Th=_ aretwostages inthepr_atadon ofcementitious materials whichshould be


_y con=oiled fortheyarecritical indeveloping theultimate resistanceofsuchmaterialto
forces ofdegradation: I)thefresh state,
during mixing, placingandconsolidation; and2)the
curing stage. Beyondthechemistry ofthecemenddous materials, factorswhichinfluence
earlystage theological propertiescan significantlyaffectthe physical propertiesof the
hardenedpasB which conuol their ultimatedurd_ty. Poor mixing, i_,_,_yate di_on,
bleedingsegregationandrelated phenotta_ cangive riseto inhomogencides inthehat,ned
paste mic_=ucmre, which can _ pathwaysforrapidtransportof harmfulspecies.
Them am _ _ limiuforwwm reductim whichpem_ adequate wodmbility.
md sula__ we ccmm_y used to improvefluidity, buttheiruse may
also gemate effem as yet. pcxx'lyundmmcaL

A combimdon of chemical and physical facu_ are influendM inthecuring stage: e.g.,
ch_..l_g thechemistry,as by pan_ sulzt_ution of Ign:mlan_ granulatedblastfunug¢ slag
or :_lim fume for cemmt can preventexcessive heat¢voludon which would remit ia
micm-cracHng. This is above and beyond theireffect in generat_g a finer pore mucture.
The chemical, physical, and mk:mstrucn_ pmperuesof hardenedcement pastes, therefore,
arecriticallyinm_elatedwith respectto theireffectson degr_n.

Poro_rv andPore Structure


ofCement Pastes

• A wideandpore
disuibudon varietyof techniques have
charact_sdcs been usedfor
ofpordand cement the d_adon Although
pastes. of themm_-my
pore size
intrusion pomsimmry(MIP)hassomelimitations" it hasbeengenerallyrecognized that the
pore_, w.hjc_.i___n_Asure.s, isrelated to the same factorswhichconlxol pm_ne.ation of
fluids• ,q_l"10115. t l'_'°'|W'_ t i ¢Gas adsorpdon methods describethe finer pore sizes of gel -
pon_t_'J butthesearenotbelievedtocontribute significandytotheu'anSlX_rateof the
variousspecies.The resultsof..MIP._nts report.edherehavebeenobtainedrlwistlh
freeze-d_jsamples treatedto numm_ damageto thespocu'nens be_'ore
measm-.,m_t.L'o_

77
Sometypical resultsof mercuryintrusionporosJmea-ystudiesare shown in figures I and
2. Plotswere rm_ of va_rlous experimen_ panmemrs, primarily: a) in_rude, d volume
(expressedascm_/cm_ providing a total porosity) asa functionof porev_dius,and b) dV/cLP
asa functionof poreradius. Surfaceareameasurementswere also made. One of the most
useful parameters hasprovedtobe thedV/dPvs._.pgr¢,,_,diusplot. fromwhichthe"critical
' poreradius" or the "maximum continuousporeradius t_)could bedetermined. Since the pore
radiusandinmasionpressurearerelatedreciprocally, the functiondV/dP may be usedinstead
of a directfunction of the poreradius. The radius at which the maximum in theplotoccurs
representsthegroupingof the largest fractionof interconnectedpore s,.w.htchin effect,
conn'olsthe n'ansmissivityof the material. The 'median'pore radiust to,t_slis also,similar.
The critical pore radiusis related to, but notidentical to Diamond's[20|'threshold radius, the
laner occurring at alarger radius.

The aboveplots contrast thefiner poresn'ucture ofacementpastecon_ninggranulated


blast-furnace slagwithapurepordandcementpaste(thecritical poreradius oftheformeris
about 1/3 of the latter). Figure 3 conn'aststhe critical pore radii of a seriesof cement pastesof
different w/c curedat different temperatures.The critical pore r_Liusincreasesby a ten-fold
factorwhen increasingthewit ratio of the slag containing materialsfrom 0.3 to 0.6, while the
total porosityincreasesby abouta factorof two (2) in similar materials. Theseresultsand
sh-'nilar
resultsfrom other workclearly indicatethat there is aneffecton the porestructurefar
beyondthat anticipated from the increased porosityalone. Furthermore,the effect differs with.
cementcomposition and type. Figure 4 gives comparisonsof mecumulative pore vommes of
portland and portland-slag mixtures.
A second r_t_scntationof pore size diswibution dam, the median pore size is the pore
radius at which 50% of the pore volume is observed in the pore size range considered. In
most c_mentitious materials rhe critical po_ radius and the median pore radius have similar
values. Pml:x_es of different cen'_ntitious compositions may be conu'asted by comparing the
porosity, median pore radius or the critical pore radius.
The mercury porosimetrymeasu_ment results for cement pastes,andthose blendedwith
flyash,u'4nulatcd slag, silica
fumeandEriequa='=(min-u-siD showsignificant dtt'ferences.
A
series ofsampleswerecuredata tempe_nn_of38"Ctheresults ofwhicharegiveninTableI.
Re!atiy.eJy.cgmpar,,ble valuestothoselisted in TableIwereobtained withspecimenscuredat
27 C.tte, =aj Except for the pure portland cement, and mixtures with fine quartz, a steady
dec'tease in medi:_n pore sizeis shownwithprolonged curingtime.
In general,the additionof mineraladmixturesto portland cementleadsto a decreasein
pore sizeor an increasein the fraction of porosityin the finer porerangeof 15nm or less.(FigA.l
The porosity may increasethrough this incorporationas in silica fume and fly ashcasesbutthe
increase inporo=tyis only_hthecluant/.ty o.
ffreer pores.The _'¢sultsan=inkeepingwiththose
foundpreviously forsimilar systeras.[lal The effect oftheincrease infiner porecontent
indiz_dy leads to an appat_nt increase in the tortuosity or interconnectivit'yof the pores.

When porosity andmean poreradius, determinedby MIP,areexpressed asa function


of
water/cementitious solid volurn¢ratio
(w/s),
aninterestingrelationship
isshown.Figure5
givesporosity dataforcementitious mixturesasa functionofw/s(vol). A familyofcurvesis
shown,whichisdifferent foreachcementitious
material.However,when themean pore
radius isplotted.
_ a.function ofw/s,(Figure
6),r,
he valuesfallon a singlelinearplot.
decreasing with dimiaishing wls.[21]

Diffusion and Permeability

As a key to the factorsinfluencingthe rate of penetntion of chloride ion into hardened


cemendrlousmaterials,and their consequenteffects, the diffusion of C1" ionsthrough
hatdenedcement pasteshas beenstudiedby a numberof investigators.[4,5] The diffusion of
CI" ionsis stronglyinfluenc_ __ by the typeof cement,and blendedcement pasteswith 30% flv
ashor 65% slaggave lower diffusion thanportland cement pastes.[5] The durability of "
blendedcement,=to highly ooncenlxatedchloride solution(27.5% C.aCI.,4.3.9% MgCI-, .,-
1.8% NaC1)is another matter, as was reportedby Feldman.[221He als_ found that bl¢'nds

78
I I I _ ] ] l i J I

0,7
CRITICAl. PORE RAOItJ$

aG
W/C * 0,40
i4 DAYS
-'-- a= 4S'C
OIFIr[RIrNTIAL

"o

O I 1 r I i I I r , [
I I I i I I i z I | (

(2J _a
2_Q _25 7'3 SO 40 30 ZS ?.0

POR[ RADIUS (ANGSTROMS ; 140* WETTING ANGLE)

Fig. 1. clV/aP plots of comparable portland cement and slag


cement pastes (60:40).

CUMULATIVE PORE VOI.UME


t l I I [ _ t [ I I
020,P" W/C • 0.40

0.1,/ ,: 0AYS CEMIrNT/SLAG..,/

?
[,_
0 F . '
I _ I 1 I I i I I I

:) CI[MI[NT --
..I

o>0.o-

0.10 _

°-;- / 0 ;..,,,¢¢ r l _ ; I I I I
• _ t
ZSQ 12S 7_ 60 _0 40 3S _ 2S ZO IS

I=OR¢"RADIUS111

Fig. 2. Cumulative Ix)re volume versus pore radius of


portland cement and slag cement (60:40) pastes.

7g
SLAG/CEMENT

_: /

'- ;
_ /
,, ....
o / ,
JO0

/
_"..." o z. II /#'
=r..' o 4_" // '/"_
....' ° _o- t/._

,o. o.,o,_ o.. o, o,


WIC RATIO

Fig. 3 Critical pore radius (dV/dP maximum) versus


w/c ratio of slag cement pastes cured at different
temperatures.

0,_ l I | /l /(

___
'_ 0.3 14 OAYS _L Ze DAYS


O CEMENT 0

2 oz
• o_,,,_
,-_ /
t o.....o%,,.,,._-o

o,- o .I/ _ _.._..,__.._t.

SLA(
W 0.04 • r W,'_
-- ! 0 o.e

AO.3 •
• 002 CEMENT SLAG
(.1

_ 0.01 I
2T
I
45
I
60
.¢," t
90
f l
27
t
45
oo.,.I
60
//" )
90

:_ CURING TEMPERATURE('_)
Fig. 4 Cumulative pore volume of portland cement
compared with slag cement pastes.

8O
I I I I

w/l?O.6vCl o ¢IFA (8012_I
4 e C/CA (70/.I0)
I wls, "¢IFA (60140)

• C/S 1351651 ,,

_, 0.4 ii//' (

Q2 [

0.I I, I , I I'
2.0 1.7S LS 1,25 t.O 0.?5
wol'er/solid (voi)

Fig. 5 Porosily as a function of w/s volume ratio;


abbre-(Hg porosimetry). C=cement,
FA=fly ash, SF=silica fume, S:,granulated
blast-furnace slag.

i i i i
%
\
- %,_\'
| -\

1.
\, \
" io- \ -_
- \
,%
I I I !
2.0 1.75 LS 1.25 1.0
waterlsolld ratio (voi)

Fig. 6 Mean pore radiusvs, w/s (vol.) ratio. .,

with 35% fly ash or 70% slag had a finer pore size disu-iburion than Type 129nland cement
and a Iowa"Ca(OH) 2 content, which improved their durability. Page et.al.t_J rcponed
chlo_'j'_le
diff_idvit7 fromsodiumchloride whichfollowed FickslawwhileKumar and
RoytOJshowedthecomparisongiveninTable2 fordifferent typesofcementpastecompared
withpordandcement(OPC),when CsClwas usedasa d/ffusam.

81
Table1. Porosity
aadMed_ Po_SizeforType! Portland
Cements
Blended
wtd_VariousAdmixtures
andCaredtt 38"C.

porosity (%_ _?edian Pore Size inrn_


wls w/s w/s

Compaddm Ap w/s w/s 0.35 0.30 w/s */s 0.35 w/s


(by v_.) (De/s) 0.40 0.35 * SP 0.40 0.35 + SP 0,30

TypeI 7 27.0 24.0 21.5 12.5 12.0 9.2


Portland 28 23.5 22.0 14.0 15.0
Cement 90 24.5 19.0 19.0 12.5
180 , 16.5 12.8 7.5 5.4
365 27.0 18.0 17.0 13.6 17.5 14.0 8.8 12.5

35%TyI_ [ PC 7 190 14.5 13.5 37


+ 65%BI_ 28 11.5 12.5 2.75 2.65
Furnace
Slag 90 10.0 6.5 2.45 2._5
180 8.0 5.7 2.40 2.55
365 6.0 4.5 2.50 2.65

90% Type[ PC 7 27.7 8.0


10%Silica 28 27.5 7.5
Fume 90 24.0 20.5 6.2 4.1
180 23.0 ,:.2
365 22,0 16.5 8.0 4.2

75%Typel PC 90 20.5 17.0 8.5 5.5


+ 12.5%5_un 180 19.5 16.5 8.5 5.0
Min-u-sil
12.5%10tin1 365 23.5 18.5 10.0 90

70%TypeI PC 7 26.6 25.0 29.0 100


+ 30%FlyAsh 28 26.0 21.3 9.2 6.-;

Min-u-sil= finepowdered
qua_ (SiO2).

Table 2
Chloride Diffusion Coefficients with CsC1 Diffusion
w/c = 0.35

Cemmt Dxl09 cm2S'I


& Blended 28 Days 90 Days

OPC 75.1 57.2


_ Slag 65 9.62 4.87
Si Fume 2.9 1.12
Fee Quartz 25 -- 67.7
Fly Ash 30 55.8 -

82
The rcvas¢ of the a-_,.sponof ions orchemical speciesinto a cement pasteis the movement
of water through the paste and remov.alof specie_i_,!eaching, which dependson the pore
su.ucmm and wat_ permeability or me paste.t',.o._._._'qAs water permeationinvolves the
u'anspor¢of species undera pressuregndiem, leachinginvolves the transportof species under
a concenwadon gradient, e.g., in the case with very purewater in contact with cement paste.
Soluble ions, particularly monovalent ones,havebeenfoundtobereleased by a diffusion
mechanism,whiletherelease rateof otherionsisgenerally slower, controlled by a
combination ofmechanisms.[ 12]

P.r,
mmlgilx
Numerousexperimental
studies
havebeenmadeoftheeffect ofwater/cementratio,
curing
temperature
andotherf
• •
actors
onthe

pgrositv,
4,
port
"_
sized istribution
• •
andpermeability
*
of
pastes
andmortars
ofcementmous matenals.[ The permeabdtty
'15.2"1.25-,8l towater
was
measured
on numerous
specimensandwascalculatedfromtheflowrate asfollows:

(viscosity
ofH20)(flowrate)
(sample
length)
Permeability
=
(cross
sectional
areaofsample)(pressure
difference
across
thesample
l

Inonestudy [4]
itwasshownthat. although total
porosities
ofcementpaste samples
curedat 60"C are smallerthan in those cured at 27"C,thepore volume larger than 750]kradius
is greater in the 60 C samples and it related to higherpermeabilitiesalso in the latter.[q Plots
of log permeability vs. w/c in Fig. ) show a nearly linearincrease with w/c.

Commonly it has been the experience that well cured materials with low w/c have low
permeabilities below the limit of measurement of many,.tzpesof apparatus.[15.21._.-31 ! This
would be expected fromoriginal predictionsof Powerst_'J and others that the "capillary"
porostty is essentially "zero"in cement paste materialswith w/c <0.40, and that the remainin,.,
porosityis fine gel pores. In one study of slag-cement pastes[15]water permeability
measurements which were made on samples taken directlyafter curing,mildly surfacedried.
and epoxied into rings, showed that Rermeabilitiesof all the 0.40 and 0.50 w/c slag cement
samples were below 10"_Darcy [I.tm':](no measurable flowunder the given test conditions)
upto60"C.

• Inother
studies
arelationship
between
water
pem_ability
andmeanporeradius
was
found.[21]Figure8 showstheapproximately
linear
relationship
foundbetween
logwater
permeability
andmeanporeradius,
forthesamematerials
asinFigs.5 and6,whichinclude
pastes
ofdifferent
w/cmadeofpureportland
cement,
andofblendswithflyash,
silica
fu,ne.
andslag.Finally,
aseries
oflinear
relationships
wasalsofoundbetweentherapid
chloride
"permeability"
(coulombs
charge
transported)
andmeanporeradius
fortypeIcementpaste
iO-S-

.IO "4-

._ IO-7 and.wlc
permeabittty
g

"o7 / l T I I I
0.35 0.40 0.45 0.50
W/¢
83
IO'W I 1 ! I I

I0"e - I
T. //

/ o
t co_' / Fig. 8 Water per_eability
i _t / vs.mean pore radius
8 ,o of pas=es from Fig. 5
m
_, .-- • •
I I I I I
iO 20 30
tte_mPorei_Ivs (nml

I..,,L,.... I I I I I I
6000 - "" """-i\
\
\
\
- \
\
\

4000- - _\\ \x (rapid chloride


Fig. 9 Coulombs charge permeabillry)
cransporl:ed
vs. water permeability, pastes
Q of Fig. 5
ZOO0- "_

I e_......._ %0
i _-- "-;_._-.._,L I
I , i I I
I 17 I
,o-' ,o
_ ,o-",40 to
_' ,o
-iio _'
water Permeability cm,s"4

andforeachofthreelevels offlyashsubsrimuon. Withineachmaterial, thechloride u"anspon


rateincreased linearly withmean IX_S,iZ_.Therewerenotenoughdataforslagandsilica
fume toestablish clear relationships,
t"JTherelationships between"rapid chloride
permeability" andwaterpe:meability followsimilar trends,
asshown inFig.9. Surprisingly,
the silica fume material exhibited higher chloride u-ansportnot consistent with the low water
permeability;, however, not enough samples of silica fume (or slag,) were investigated to make
validcomparisons withtheotherblended ruaterlab. Additionalstudies arerequired tovalidate
these relations for all types of materials.

and Conclusions
Discussion

As discussed above,permeabilities towateranddiffusion ofionic species in


cementitious grams,pastes andmortars areimportant keystoconcrete durability. Therefore
investigations have been made of numerous materials containing portland and blended
cements, and those with fine-grained admixtures (fly ash, slag, silica fume), at room
temperatureandafter prolongedcuringat severalelevatedtemperaturesup to 90"C These
constitutepart of studiesof fundamentalmaterialrelationshipsperformedin order to address
the question of long-termdurability. In zeneral,the permeabilitiesof the materialshavebeen
' found to be low [many < 10"8Darcy ( I0"'13m •s"I)] after curing for 28 days or longer at
temperatures up to 60"C. The results obtained at 90"C are somewhat more complex.

: Porosity, mean poreradiusandporesizedistributionin cementitiousmaterialsarevery


important micro-structuralcharacteristics;theyarerelated to a seriesof properuesof materials,
suchas flexural sta'ength,fracture toughness,durability, and resistivity to ionic diffusion.
The work discussedhere revealsthefactthatthe poresmactureof materials in questionis
essentiallyaffected by the fineness,w/s ratio,chemicalcomposition,andpozzolanicreaction.
Chloride ion diffusion under an appliedelectricfield appearsto be affectednotonly by the
pore smactu_ of the materials, butalsoby thediffusion mechanism.
84
The relationship between porosity, pore structure and ionic diffusion was suggested as
follows:

cc= [Ca rCb]3/e

whereCa andCb areconstants;e isporosity,


rismedianporeradius,
ctffi
ratiooftheoretical
to
effectivediffusion.[
6.16]ValuesforCa andCb weregenerated fortheneatportland cement
oastesand C, = I03 + 2.6andCb ffi
-0.41± 0.13forCI"diffusion
andC a = 4.6± 2.5andCb
= -0.62± 0.0_forCs wereobtained overthetemperature
rangeof27 to60 C.

A relationship hasbeenfoundbetweenmean poreradius and"rapidchloride


permeability,"
aswellaswaterpermeability. Additional studies arerequired torefine
theserelationships
and
determine their applicability to a variety of ceracnridous materials.

Acknowledgement

The research described herein was supported by the Su'ategic Highway Research Program
(SHRP). SHRP is a unit of the National Research Council that was authorized by section 128 of
the Surface Trauspormtion and Uniform Relocation Assistance Act of 1987.

REFERENCES

1. P.K. Mehta and D. Manmohan, Prec. 6th Int. Congr. Chemistry of Cement, Vol.
HI, Theme VH, 1-5, Editions Septima, Paris (1980).

2. B.K. Nyame and J.M. Illston, Proc. 6th Int. Congr. Chemisu'y of Cement. Vol. III,
Theme VI, 181-185, Editions Septim& Paris (1980).

3. H. Ushiyama and S. Goto, Proc. 6th Int. Congr. Chemistry of Cement, Moscow,
2(1), 331-337, Su'oyizdat, Moscow (1974).

4. S. Goto and D.M. Roy, Cem. Concr. Res. 11(5), 751-757 (1981).

5. C.L Page, N.R. Short, and A. El Tarars, Cem. Concr. Res. 11, 395-406 (1981).

6. A. Kumar and D.M.Roy, Proc. 8th Intl. Congr. Chemistry of Cement, Brazil, V.V,
73-79 (1986).

7. R.Parsons,
HandbookofElectrochemical
Constants.
Butterwonhs
Scientific
Publications,
Table79,85 (1959).

8. H.G.MidgleyandJ.M.Illston,
Cem. Concr.Res.14.453-614(1984).

9. A.AtkinsonandA.K.Nickerson,
AERE Harwell,
DOE ReportNo.DOEARW/83,
137pp.

I0. K.Anderson,B.Torstenfelt.
andB.Allard,Scientific
BasisforNuclearWaste
Management3.235-242,Ed.J.G.Moore,PlenumPress,
NY (1981).

lI. R.Heitanen,
T.Jaakola,
andJ.K.Miettinen,
8thInt.
Syrup.Scientific
Basisfor
NuclearWasteManagement,Boston,MA (Nov.26-30,1984).

12. D.M. Roy,Proc.8thIntl.


Congr.Chem.Cement.Brazil,
V.I,362-380(1986).

13. D.M. Roy,R.Malek,andP.H.Licasu'o.


ConcreteDurabiliw.
K.andB. Mather
International
Conference,
AC[ SP-I00,Vol.2,1459-1476(i987).

85
14. D.M. Roy and R.I.A. Malek, Proc., Intl. Workshopon Granulated Blast Furnace
Slag in Concrete, Missasauga,OnL(1987), 12pp.

15. D.M. Roy and K.M. Parker,Proc.CANMET/ACIFirst Intl. Conf. on the Use of
Fly Ash, Silica Fume, Slag and Other MineralBy-products in Concrete, Vol. l, Ed.
V.M. Malhoa'a,pp. 397-414; ACI SP-79, ACI, Detroit (1983).
16. A. Kumar, Diffusion and PoreStructureStudiesin Cementitious Materials,Ph.D.
Thesis in Solid State Science, The PennsylvaniaState University, University Park.
PA 16802 (1985).

17. R.F. Feldman, Theme 4.1, Proc. 8th Intl. Congr.Chem. Cement, Brazil, V. I.

18. A. Kumar and D.M.Roy, Cem. Concr. Res. _ 74-78 (1986).

19. R.I.A. Malek, D.M. Roy, and Y. Fang, Port Structure,Permeability, and Chloride
Diffusion in Fly Ash and Slag Containing Pastes and Mortars(this Symposium in
press).

20. S. Diamond, (1973). Port Structureof HardenedCement Paste as Influenced by


Hydration Temperature,pgre Structureand Pro_oerties
of Materials,Proc. Intl.
Syrup. Prague, Sept. 1973, I-B73-B88.
21. S. Li and D.M. Roy, Investigationof RelationsBetween Porosity. PoreStructure.
andCI Diffusion of Fly Ash and BlendedCementPastes:Cem. Concr. Res..L.6.,
749-759 (1986).

22. R.F. Feldman, Proc. 5th Intl. Syrup. Monterey,1981, 261-288.

23. E.L Whim, B.E. Sch¢¢tz, D.M. Roy, K.G. Ziammrman, and M.W. Grutzo:k,pp.
47-4,78in Scientific Basis for NuclearWasteManafen'_nt_Vol. 1; Proc., Materials
Rese,a_h Society, Ed., G.J. McCarthy,Plenum, NY (1979).
24. M.W. Barnes, D.M. Roy, and C.A. Langton,_i:ientific Basis for NuclearWaste
VIII, pp. 865-874, Eds., C.M..Iantzen, J.A. Stone, and R.C. Ewing
(1985); Materials ResearchSociety Symposium Proceedings, Vol. 4.4.

25. B.E. Scheetz, E.L. White, D. Wolfe-Confer,and D.M. Roy, 7th Intl. Congress
Chemistry of Cement, Pans (1980), Vol. Ill, Communciations, VI-170-VI-175.

26. M.W. Grutzock. B.E. Sch_t2, E.L White, and D.M. Roy, Borehold and Shaft
Plugging Proceedings OECDAJSDOE,Columbus,OH (7-9 May 1980), 353-368,
OECD, Paris, France (1980).

27. LD, Wakel¢y and DM. Roy, A Method forT_dng the PermeabilityBetween Grout
and Rock. Cera. Concr. Res. L2,,533-534 (1982).

28. D.M. Roy, E.L. White, and Z. Nakagawa, Effects of Early Heat of Hydrationand
F,xposunt to ElevatedTemperatureson Pro[:smiesof Mortarsand Pastes with Slag
Cement, ASTM, STP, 858, Tcm_rature Effectson Concrete. 150-167 (1985).
29. T.C. Powers, J. Am. Ceram. SOC._ 1-6 (1958).
j.

30. B.E. Sch¢*tz and D.M. Roy, pp.933-942 in Scientific Basis forNuclear Waste
M.gOggtllIlt_,Vol. VIII; Proc., MaterialsResearchSociety, Vol. 44, Eds., C.M.
._ Jantzen, J.A. Stone and R.C. Ewing (1985).
31. B.E. Scheetz, D.M. Roy, and C. Duffy (in press).

86
Concrete and Structures Advisory Committee

Chairman Liaisons
James J. Murphy
New York State Department of Transportation Theodore 1L Ferragut r
Federal Highway Administration
Vice Chairman
Howard H. Newlon, Jr. Crawford F. Jencks
Virginia Transportation Research Council (retired) Transportation Research Board "

Members Bryant Mather


USAE Waterways Eaperiment Station
Charles J. Arnold
Michigan Department of Transportation Thomas J. Pasko, Jr.
Federal Highway Administration
Donald E. Beuerlein
Koss Construction Co. John L. Rice
Federal Aviation Administration
Bernard C. Brown
Iowa Department of Transportation Suneel Vanikar
Federal Highway Administration
Richard D. Gaynor
National Aggregates Association/rNational Ready Mixed 11/19/92
Concrete Association
Expert Task Group
Robert J. Girard
Missouri Highway and Transportation Department Amir Hanna

David L. Gress Transportation Research Board

University of New Hampshire Richard H. Howe

Gary Lee Hoffman Pennsylvania Department of Transportation (retired)


Pennsylvania Department of Transportation Stephen Forster

Brian B. Hope Federal Highway Administration


Queens University Rebecca S. McDaniel
Indiana Department of Transportation
Carl E. Locke, Jr.
University of Kansas Howard H. Newlon, Jr.

Clellon L. Love.all Virginia Transportation Research Council (retired)

Tennessee Department of Transportation Celik H. Ozyildirim

David G. Manning Virginia Transportation Research Council


Ontario Ministry of Transportation Jan P. Skalny
W.R. Grace and Company (retired)
Robert G. Packard
Portland Cement Association A. Haleem Tahir
American Association of State Highway and Transportation
James E. Roberts Officials
California Department of Transportation
Lillian Wakeley
John M. Scanlon, Jr. USAE Waterways Experiment Station
Wiss lanney Elsmer Associates

Charles F. Seholer
Purdue University

Lawrence L. Smith
Florida Department of Transportation

John 1L Strada
Washington Department of Transportation (retired)

You might also like