Sie sind auf Seite 1von 43

Research and Development Laboratories of the Portland Cement Association

RESEARCH

DEPARTMENT

BULLETIN 80

Some Aspects of the Physics and Chemistry of Cement

BY STEPHEN BRUNAUER

AuthorizedReprint from THn SCXENCSI OFENCiXNEERINQ MATERXALE!, Edited by J. E. Goldman John Wiley & Sons, Inc., New York Copyright 1957

Research and Development Laboratories of the Portland Cement Association

RESEARCH

DEPARTMENT

BULLETIN 80

Some Aspects of the Physics and Chemistry of Cement

BY STEPHEN BRUNAUER

AuthorizedReprint from THESCIENCE or ENGINEERING MATERIALS, Edited by J, E. Goldman John Wiley & Sons, Inc., Ncw York Copyright 1957

16

Some

Aspects

of the Physics

and C~en)istf=yof Cement

STEPHEN

BRUNAUER *

INTRODUCTION The subject for this chapter is the physics and chemistry of cement. That is, I am afraid, a big order for one chapter. Many long books have been written on the subject, and, naturally, none of them, nor all of them together, cover the entire field. Thus, selection must enter which implies personal bias, The author selects what he considers most important or most interesting or, simply, what he knows most about. I am relatively new in this field; at the time of writing I had been in it for only two and a half years. If Dr. Goldman had selected someone else, the ~ubjects covered would have been considerably different from those selected by me. Only a few names appearand those with considerable frequency. Thousands of able scientists have been working in the cement field whose names do not appear in this chapterindeed, some of the greatest names are missing. By far the largest part of the work reported here was done by the research scientists of the Portland Cement As*Principal Research Chemist, Portland Cement Association Research and Development Laboratories. 428

Physics and Chemistry oj Cement

429

sociation. These facts require some explanation. The emphasis in this chapter is on that part of basic physiochemical research which has revealed to us most about the properties of those engineering materials whose most important constituent is portland cement. With this in mind, the choice is as good a one as could be made. Nothing in this chapter would delight the eyes and ears of the solid-state scientist-no wave equations, no energy bands, no exclusion principles. Cement is much too complicated a system to be attacked with the theoretical tools of solid-state science at this stage. What is reported here is physical chemistry and not solid-state physics; it is basic research, but not fundamental research. Even so, it will be of some value for the engineering educator in enabling him to appreciate that there is even some science in so mundane a structural material as cement. Research in the field of the basic chemistry and physics of cement involves the investigation of three systems: (1) cement; (2) cementwater; (3) cement-water-aggregate. The aggregate is sand, gravel, crushed stone, or some similar substance. The two most important industrial materials that belong to the third system are concrete and mortar. The three systems may be loosely referred to as a one-component, a two-component, and a three-component system. The reference is loose because portland cement itself is, strictly speaking, a manycomponent system; it is a physical mixture of several distinct and not too difficultly identifiable chemical compounds. Nevertheless, portkmd cement in its reaction with water behaves, to a considerable extent, like a single compound, and in the hydration product the separate hydrates of the initial cement compounds cannot be identified. Perhaps it would be more accurate to say that separate hydrates have not been found to date; it is very likely that future research will detect the separate hydrates, if they exist. Cement is mixed with water in such proportion as to form a plastic mass, This mixture, called the cement paste, in a few hours loses its plasticity and becomes comparatively rigid. This is called the setting process. After setting, the cement paste continues to harden for months, possibly even for years, and it becomes in appearance and texture like a stonethis is the hardening process. The hardening process is, at least in part, a chemical process of hydration. Some changes in the hardened paste occur even after hydration ceases; this is sometimes referred to as the aging process. Concrete is an artificial rock, produced by the interaction of cement,

430

The Science oj Engineering Materials

water, and aggregate. Cement and water reack to form the hardened paste, which is the matrix that binds the masses of aggregate together, The main structural properties of concrete are determined, to a considerable extent, by the properties of hardened paste. Although the presence of the aggregate contributes in an important way to the properties of concrete, an understanding of the behavior of hardened paste gives a first approximation to the understanding of the behavior of concrete, The bulk of this chapter deals with the two-component system, cement-water, and especially with the physical structure of hardened portland-cement paste. The one-component system is discussed only briefly, and the three-component system is not discussed separately, at all. Whatever is said about concrete appears as a part of the discussion of the two-component system. It is difficult to cover such a broad field adequately in a single chapter, even in as long a one as this; it is more instructive to spend most of the time on one central theme rather than to attempt to discuss many. PORTLAND CEMENT

Cement is as old as human civilization, or older. The pyramids of Egypt are among the most ancient structures erected by man; between their gigantic blocks of stone we find cementing material. The cement used by the Egyptians was a calcined, impure gypsum; the Greeks and Remans used calcined limestone.1 A very large fraction of the cement used in our age is a manufactured product called portland cement. (Last year more than 60 million tons of portland cement were produced and sold in the United States alone.) Portland cement is made by intimately mixing a calcareous (lime-containing) and an argillaceous (clay-containing) material; heating the mix at a high temperature, up to 1450 C, usually in a rotary kiln; cooling the resulting clinker; mixing it with a few percent gypsum; and grinding the mixture to a very fine powder. The order of magnitude of the average dimension of portland-cement grains is 10 microns (0.01 mm). The chemical reactions occuring in the kiln are not discussed here but the chemical nature of the clinker is analyzed. Extensive research, including chemical analyses, phase equilibrium studies, microscopical, optical, X-ray diffraction, and other types of investigations, has established that portland-cement clinker is a mixture of several chemical compounds. Four of them are usually regarded as the major constituents;

Physics and Chemistry oj Cement

431

they are two calcium silicates, a calcium aluminate, and a calcium aluminoferrite. The minor constituents are uncombined calcium oxide, magnesium oxide, and small quantities of other impurities. Chemical analysis also shows the presence of small amounts of the alkalies NazO and KzO. In addition to the constituents of the clinker, portland cement contains a few percent gypsum, as stated above. Since the cement compounds are repeatedly referred to in the course of the discussion, to avoid cumbersome long words and chemical formulas, we shall introduce the following symbols used by the cement chemists: CaO = C, Si02 = h, A1203 = A, FezOs = F, KzO = K, Na20 = N and HZO = H. The four major constituents are CWS= 3Ca0. SiOz, tricalcium silicate; C2A= 2Ca0. Si02, dicalcium silicate; C3A = 3Ca0 oA1203, tricalcium aluminate; and CAAF = 4Ca0. AlzOa. Fe203, tetracalcium aluminoferrite. Actually, the phase that contains the iron oxide is a solid solution, ranging in composition from C2F to COAZF, as was shown by Swayze? C4AF k a member of this solid-solution series, and, for simplicity, we may take it as representative of the iron-containing phase. Among the five principal types of portland cement in this country, by far the most widely used is Type I or normal portland cement. The composition of a representative Type I cement is shown in the following table.
cd C& (73A C4AF

~= 4570 = 27~o
= Ilyo = 8%

CaSOl MgO

= 3% = 3yo Cao = 0.57* Na20, KZO = 0.570

The other four types contain the same constituents in somewhat different proportions. Our knowledge of the computation of the quantitative compositions of portland-cement clinkers is based on the fundamental work of Bogue,S In order to understand why we have different types of portland cement, we must know something about the reaction between cement and water. In the hydration of cement a considerable amount of heat is evolved; roughly 120 calories per gram for a Type I cement. In large concrete structures, such as dams, for example, in which the heat of hydration is not readily dissipated, undesirable thermal stresses may be introduced. In order to guard against such deleterious effects on the structure, cements that evolve less heat were developed by altering the composition of the cement. It was found that C3A hydrates very rapidly, with large heat evolution; C& hydrates slowly, with

432

The Science of Engineering Materials

much smaller heat evolution; and C3S is intermediate between them, both in the rate of heat evolution and in the quantity of heat evolved. Before the building of Hoover Dam a new cement was developed, now known as Type IV or low-heat cement, in which the percentage of C2S was increased, and that of C3A and CW!3 was diminished. Type II, or moderate-heat-of-hardening cement is intermediate in composition between Type I and Type IV. Large heat evolution is not always undesirable. At freezing temperatures, for example, cement reacts with water sluggishly and with ice not at all. The heat evolved in the reaction keeps the system at higher than ambient temperatures and provides for the continuation of the reaction. Type 111 cement evolves more heat in its hydration and faster than the normal Type I portland cement. This is a cement high in CJ3 and C3A, and low in CJ3; it is also ground finer than Type I cement. The rapid hydration is accompanied by rapid hardening and strength development; this cement, therefore, is called highearly-strength cement. Because of this characteristic of Type 111 portland cement, it finds application also in concrete structures, when rapid strength development is desired. For example, in the building of concrete highways Type III cement is used when there is a special urgency to open up the traffic on the road as soon as possible. Type V, or sulfate-resisting cement is not discussed. The foregoing brief discussion of the various types of portland cement simply illustrates how basic chemistry influences important engineering properties of the substance. The differences in chemical composition are achieved by using different proportions of the various raw materials for the kiln feed. The four major components of portland cement are what the chemist calls calcium salts. The calcareous or calcium-supplying raw materials are limestone, cement rock, chalk, marl, marine shells, and alkali waste. The acid parts of the salts, which contain the silica, alumina and ferric oxide, are supplied by the argillaceous materials: clay, shale, slate, blast-furnace slag, ashes, and cement rock.1 In the portland-cement industry great care is taken in the proportioning of the raw materials to obtain each type of cement within the specified composition limits. Important contributions to the theory of proportioning were made by Dahl.4 The crystal structures of the major constituents of portland cement have been extensively investigated. Since all of them are complex crystals, with large unit cells and low symmetry, the structure of none is closest to soluof them is completely solved. The structure of CsLS

Physics and Chemistry oj Cement

433

tion; it was published by Jeffery .5 C2tS exists in four crystalline forms, designated as a, a, /3, and y; but the structure of none of them is fully worked out.8 Portland cement contains predominantly the /3 form; it is uncertain to what extent the others appear in it. The crystal structure of C3A was discussed by 0rdway,7 who has been working on it for some years. Least is known about the structure of C4AF and the other members of the iron-containing solid-solution phase. The problem is still further complicated by the fact that in portland cement none of these crystals appears in pure form. The Call phase contains some dissolved aluminum and magnesium, the CZS phase contains potassium oxide, and the C3A phase contains sodium oxide. The iron-bearing phase has a variable composition, with differing proportions of CaO, A120~, and Fe203 in it. In addition, because of rapid cooling, cement contains a varying percentage of glass (2 to 20 percent in commercial clinkers) ; 8 and part of the (73A and C4AF (frequently the greater part) is in this glass phase, and not in crystalline form. The above facts show that portland cement is a complex physical and chemical system. The cement-water system is, of necessity, even more complex, since it includes not only the reactants, i.e., water and the compounds in cement, but also the reaction products, i.e., the hydrates formed in the reactions. If, in order to obtain some understanding of the cement-water system, we should first have to know the behavior of each component separately and then synthesize a physiochemical model from the information gathered, our knowledge of the system would be in a very rudimentary stage, indeed. Fortunately, this is not true. A considerable amount of basic information has been obtained by treating portland cement as a single component, as we shall see. THE CEMENT-WATER SYSTEM

An idea of the fineness of portland cement can be gained from the fact that 1 pound of it consists of about 500 billion particles, If a pound of cement is mixed with half a pound of water, after a few hours the cement paste becomes hard as a rock, and to all outward appearances it looks like a homogeneous, solid mass. However, appearances are deceptive; the hardened paste actually contains matter in all three states of aggregation: solid, liquid, and gas. The rock is porous; roughly one-third of the volume of the hardened paste is pore space, and the pores are filled with water and air. The solid phase

434

The Science oj Engineering Materials

itself is not a single mass; it contains matter in a very fine state of subdivision, in the colloidal state. The order of magnitude of the particle dimensions produced in the hydration of cement is 100A; thus in the hydration process each of the extremely fine grains of cement powder gives rise to a billion particles of hydrated cement. These particles, whose diameter is only one order of magnitude greater than the diameter of the molecules, together form a continuous, porous, solid mass, called the cement gel. The investigation of the cement-water system may be divided into two parts: (a) investigation of the hydration process; and (b) investigation of the product of the interaction of cement and water, i.e., cement paste. Logically (a) should be discussed first, since the process precedes the product. However, for convenience, (b) will be discussed first. The physical changes accompanying the hydration of cement were discussed in the introduction. On the basis of these changes, the investigation of portland-cement paste may be divided into the following stages: (a) the plastic paste; (b) structural changes occurring in the setting process; (c) the hardened paste; and (d) structural changes occurring during the aging of hardened paste. Of the four subjects only the hardened paste will be discussed. Because understanding of this subject is more important than that of the others, far more work has been done on it than on the others. Nevertheless, interesting and important work has also been done on the other subj ects.O
Hardened Portland. Cement Paste

Hardened cement paste is a truly remarkable substance, produced by the collaboration of man and nature. The first important contributions to the understanding of the physical structure of hardened paste were made more than 60 years ago by Le Chittelier,l with his crystallization theory, and Michaelis with his colloid theory. There is no room left for doubt now that Michaelis was fundamentally right hardened paste is a colloidal systembut in some of the details Le Chiltelicr was closer to the truth than Michaelis. Their theories gave a qualitative picture of the process of cement hardening and of the physical structure of hardened paste, and we may call this pictureto use a favorite term of the theoretical physiciststhe zeroth approximation. In the subsequent decades many investigators added their contributions to this first, qualitative picturesome supporting Le Chiitelier, others Michaelis, and still others advocating a com-

Physics and Chemistry

of Cement

435

promise between the two views. All this work may be regarded as a modification, refinement, and improvement of the zeroth approximation model. Since our concern is more with present-day understanding than with history, we shall not discuss either of these theories. The beginnings of a new approach to the study of hardened paste in 1931.18 were made by Jesser lZ in 1929 and by Giertz-Hedstrom Stimulated by their work and by a hypothesis advanced by Freysinnet 14in 1933, Powers and his co-workers carried out extensive experiments that resulted in the first thoroughly quantitative model of the physical structure of hardened paste, In 1946 and 1947 Powers and Brownyard 15published nine papers in the Journal of the Arnemcan Concrete Institute, with the title, Studies of the Physical Properties of Hardened Portland Cement Paste. These papers were collected and published together as Bdletin .2% of the Research Laboratories of the Portland Cement Association. Most of this chapter is based on Bulletin .%?and on subsequent work of Powers and his co-workers. One logical way to proceed would be to divide the discussion of hardened paste into three parts: (a) The solid phase L Physical structure 2, Chemical constitution (b) The liquid phase 1. Properties of adsorbed water 2. Properties of capillary water 3. Effects of dissolved substances (c) The gas phase 1. Effects of water vapor 2, Effects of entrained air 3, Effects of carbon dioxide Although it would be most satisfactory to follow such an outline, in practice it is almost impossible to do so. The various aspects of the problem are so closely intertwined that most of the experimental attacks furnish information about more than one phase, simultaneously. Take, for example, adsorption experiments. Investigation of the adsorption of water on hardened paste contributes valuable information to five of the eight above subjects. As a matter of fact, adsorption studies have proved to be about the most valuable tool employed to date in the study of the structure of hardened portland-cement paste. This technique was introduced into the field by Powers, and a considerable portion of Bulletin .%?

436

The Science oj Engineering

Materials

deals with adsorption studies. The discussion of the structure of hardened paste will start with these water-vapor adsorption studies. Water adsorption. Adsorption is the retention of foreign molecules on the surface of a substance. Wherever a surface exists, there is also adsorptionthe surface of every solid and liquid attracts every molecule, regardless of its chemical nature. Because of this lack of chemical specificity, this type of adsorption is called physical adsorption; and because the forces operative in the binding of the adsorbed layer are predominantly van der Waals forces (see Chapter 2), it is also The specificity is not entirely called van der Waak adsorption. lacking; the force of attraction does depend to some extent on chemical factors, such as the atomic or ionic structure of the solid and the chemical nature of the molecule in the vicinity of the surface-but only to a slight extent. The dominant factors are physical: temperature, pressure, and concentration, Adsorption of a vapor on a free and uniform surface depends upon three factors: the extent of the surface area, the energy of interaction between the vapor molecules and the surface, and the energy of interaction between the vapor molecules themselves. In principle, these three factors can be evaluated from adsorption isotherms, i.e., from measurements of the quantities of vapor adsorbed at various pressures, at a constant temperature. The theory most widely used at present for the interpretation of physical adsorption is the Brunauer-EmmettTeller theory6 (see Chapter 11). This theory gives only a rough approximation of the energy factors involved in the adsorption process, but it gives the extent of the surface area with considerable accuracy. When adsorption takes place in the pores of an adsorbent, and not on a free surface, the amount of vapor that can be adsorbed is limited by the sizes of the pores. This introduces a fourth factor, namely, the width or diameter of the pore. At the saturation pressure of the vapor, p8, all the pores of the solid are filled; consequently, the water uptake at saturation gives us the total porosity of the solicl. In addition, the shape of the adsorption isotherm gives us some clue as to the distribution of the pore sizes in solids. Thus, adsorption studies can furnish us information about such vitally important properties of a finely divided, colloidal system as surface area and porosity. In hardened portland-cement paste, water appears in all three states of aggregation: as part of the solid, as part of the gas phase, and as the liquid phase. The water, which is in chemical combination with the solid, has been called by different investigators fwater of constitution,

Physics and Chemistry

of Cement

437

(combined water, %vater of hydration, (fixed water, (bound water,) and (non-evaporable water. Before an adsorption experiment can be conducted on cement paste, the surface of the adsorbing solid must be first freed of adsorbed matter, the pores must be emptied, the evaporablc water must be removed.

P /P,

Fig. 16.1. Water-vapor adsorption isotherms on a hardened porthmd-cement paste at different ages. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute.) This is usually done by combining evacuation and the application of some drying agent to establish a very low pressure of water vapor in the system. Powers and Brownyard used Mg (CI04) z2Hz0, which gives a pressure of 8 x 10-3 mm of mercury. They assumed that the water remaining in the solid, or non-evaporable water, w~, approximately corresponded to the water of constitution. Some examples of the water-vapor adsorption isotherms obtained by Powers and Brownyard 1s are shown in Fig. 16.1. The weight of the water adsorbed per gram of dry paste is plottecl against the relative pressure, p/p, (p is the equilibrium pressure of water vapor in the adsorption system, and p, is the saturation pressure of the vapor at the

438

The Science oj Engineering Materials

The isotherms were determined at room temgiven temperature). perature; the four isotherms refer to four different pastes. As the relative pressure (or relative humidity) increases, the amount of water adsorbed increasesand the curves have a curious shape. For reasons best known to themselves, the adsorption chemists used 0

161

High - vacuum method

1/

8 6 4

-0.031

0.3 0.4 (5 0.2 ,x (= PIPS) Fig. 16.2. B.E.T. plots of water-vapor adsorption isotherms on hardened portIand-cement pastes. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute.) Y 0 -0.026 0.1 to call these S-shaped isotherms, until 1940 when the author christened them Type II isotherms, At any rate, practically all vapors adsorbed on practically all absorbents give isotherms of this shape, According to the B.E.T. theory, if a certain function of w, the weight of the adsorbed substance, and Z, the relative pressure, is plotted against z (or p/pe), a straight line should be obtained. Figure 16.2 shows B.E.T. plots of adsorption isotherms of water on two hardened portland-cement pastes, and, clearly, within a certain range of relative pressures, good straight lines were obtained. From the slope

k
J

Cement 13723

Physics and Chemistry

oj Cement

439

and intercept of each straight line one can obtain two parameters of the B.E.T. theory, C and V~, shown in the figure, from which can be obtained the approximate interaction energy between surface and water and the accurate surface area of the paste, respectively. Actually, V~ is the weight of water required to cover the entire surface of the paste with a single layer of adsorbed molecules. We could obtain tile surface area of the paste within a few percent, if we knew the average area occupied by a water molecule on the surface, We do not know this; consequently, we must assume an area for the molecule. The assumption, confirmed indirectly in a variety of ways, is that the packing of the adsorbed molecules is the same as that of the molecules in the liquid state. Using 10.6 A2 as the area of an adsorbed water molecule, Powers and Brownyard obtained for the surface area of an almost fully hydrated cement the value of 2,000,000 cm2/gram, or 200 m2/gram. This is a tremendous surface area, indicating that the material is very finely divided. Adsorption cannot supply information about the shapes and sizes of the particles; it gives us only the surface area. However, if we know the shapes of the particles or assume a shape, we can calculate an average particle size. Powers and Brownyard assumed that the particles were spherical in shape and thus obtained an average particle diameter of 140A. If the particles were not spheres, at least one of their dimensions would have to be less than 140A. This settled once for all whether hydrated cement was a colloid, since, by definition, a colloid is any substance that has at least one dimension less than 1000 A. The old controversy has been settled; hardened paste was found to be a colloidal system. Swerdlow and Heckman 18at the National Bureau of Standards obtained electron photomicrographs of hydrated cement, one of which is shown in Fig, 16.3. The particles were found to be approximately spherical in shape; and they ranged in size from about 50 to 250A. The conclusion of Powers has been confirmed. Knowledge of the surface area of hardened paste was a great step forward. Many of the important properties of paste are determined predominantly, or at least in part, by the fact that it is a colloid. The unique properties of a colloid are due to its large specific surface area and the large energy residing in the surface. Probably the most important basic physical property of hardened paste is its surface area. Powers and his co-workers were able to show remarkable correlations between surface area and such vitally important properties as

440

The Science oj Engineering Materials

strength, shrinking and swelling, permeability to water, and freezing and thawing. Some of these correlations will be discussed later. Figure 16.4 represents plots of total water content of hydrated

I/1.l

Fig. 16.3. Electron microgmph of a Type I portland cement fully hydrated in a ball mill. (From Swerdlow, McMurdie, and Heckman. By courte~ of the authors, the Reinhold Publishing Corporation, and the Royal Microscopical Society.) cement against relative pressure, p/ps, at different (ages, obtained Age is the term used by the cement by Powers and Brownyard.5 chemist to denote the period of hydration, The same cement was used in these experiments, and it was mixed with the same amount of water; the weight ratio of water to cement, wo/c, was 0.309. The six isotherms were obtained after periods of hydration ranging from 7 days to 1 year.

Physics and Chemistry of Cement

441

The total water is the sum of the non-evaporable water, or water of constitution, wn, and the evaporable water, or adsorbed water, we. As mentioned before, prior to adsorption the cement paste was equilibrated at the partial pressure of water vapor over Mg (CI04) ~2H20, I 0.36 0.32 0.28 E E 8 0.24 % E s w 0.20 g z s 0.16 ml j $ ().12 ~ c 0.08 0.04 00 02 04 06 I I

+365 days ~ 90 days ~ 28 days w 14 days w 7 days

0.8

1.0

p /p~at 25C Fig. 16.4. Total water isotherms on a hardened portland-cement paste at different ages. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute,) which is about 8 x 10 3 mm of mercury, The water remaining in the paste at this low pressure is Wn,and on the scale of Fig. 16.4 it appears as the zero pressure value. Please note that w. increases with time it is a measure of the extent of hydration of portland cement. After 1 year it is somewhat more than twice as much as after 7 days for this particular cement and wO/c ratio, The cement was not fully hydrated even after 1 year; it was, perhaps, about 80 percent hydrated. The curves beyond the zero pressure value represent the adsorption

442

The Science of Engineering Materials

isotherms. The isotherms shown in Fig. 16.1 were obtained from Fig. 16.4, by subtracting WW from the total water. With the help of B.E.T. plots (the type shown in Fig. 16.2), one can obtain V~, the parameter representing the surface area, for each of the isotherms. Powers and Brownyard found that the surface area increased with age, which is what one would expect, since the cement gel formed in the hydration is much more finely divided than unhydrated cement. However, they found something else that was wholIy unexpectedthe surface area increased in direct proportion to the water of hydration
v% = kwrj

(16.1)

They found this to be true for other cements and for other we/c ratios as well. Later, when the hydration process is discussed, the implications of this remarkable observation are enumerated. At this point it is legitimate to raise such questions as these: (a) How close does the W. obtained by Powers and Brownyard come to the true water of hydration? (b) How close does their V~ come to the true surface area of hardened portland-cement paste? On the basis of the work of Copeland and Hayes, those at the Portland Cement Association who do basic research on cement hydration have adopted a different method of drying the paste, The paste is brought to equilibrium with water vapor at a pressure of 5 x 10 4 mm of mercury, the pressure being produced by ice at 78 C, the temperature of solid COZ. This pressure is considerably lower than that given by Mg (C104) z2HZ0. As a result, we obtain wk values that are 8 to 10 percent lower than those of Powers and Brownyard, and V~ values that are about 30 percent higher. The reason for this is that hydrates in the colloidal state do not have sharply defined dissociation pressures; at sufficiently low pressures they give up water continuously as the pressure is diminished. Because the energy of binding of combined water in the hydrate is of the same order of magnitude as the energy of binding of the adsorbed water to the surface, a sharp and quantitative separation of the water of hydration and adsorbed water is impossible. The question is then, how do we know whether we are even close to the correct values of W. and V~? Copeland found an answer to this question. In the range x (or P/P,) = 0.05 to 0.35 (sometimes in a somewhat narrower range), the B.E,T. relation (Eq. 11.5) is accurately obeyed by most adsorption systems. He assumed, therefore, that the course of the adsorption is correctly described by the B ,E.T, relation in this range. However, to

Physics and Chemistry

oj Cement

443

obtain V~ and C from the adsorption data, one must know w, the amount of adsorbed water (see Fig. 16.2), and this is just what we do not know accurately. Copeland, therefore, differentiated the B.E.T. equation and calculated Vw and C not from the adsorption isotherm, i.e., the dependence of w on $, but from the slopes of the isotherm, i.e., the dependence of dw/dx on x. This work is not yet complete, but the indication to date is that the correct value of W. falls between the values obtained by the two different methods of drying, but closer to the newer method (ice at 78 C), than to the older (magnesium perchlorate), Using Copelands assumption but a somewhat different approach to the problem, the author also found that the correct value of W. falls somewhere between the values obtained by the two methods of drying. The uncertainties in V~ and W. discussed above have no essential influence on Eq. 16.1, For the two different methods of drying we obtain different k values, but the equation holds for both cases. Likewise, all the important findings of Powers, which have been or will be discussed in this chapter are only slightly affected by the method of drying, or not at all. Porosity and capillary adsorption. So far one very important characteristic of hardened paste, its surface area, has been discussed. Another almost equally important property of the paste, about which very little has as yet been said, is its porosity. Many of the important structural properties of hardened portland-cement paste are determined to a great extent by the surface area and the porosity of the paste, Figure 16.1 represents the adsorption isotherm parts of the curves of Fig. 16.4, Adsorption increases with ageexcept at the right ends of the isotherms, where the reverse is true. Close to saturation the curves cross each other, and the least hydrated paste holds the most evaporable water. Powers and Brownyard5 gave a clear explanation of these curious results in terms of surface area and porosity. Adsorption takes place in the pores of the paste, and as p/p, increases larger and larger pores become filled with water. At the saturation pressure all the pores are filled; consequently, the water content at p/p8 = 1 gives a measure of the total porosity of the paste. AS Fig. 16.1 shows, the porosity decreases with age, Powers explanation is the following. When the plastic paste sets a few hours after mixing, it forms a rigid framework which fixes the total volume of the hardened paste permanently. If the total volume is constant, it is clear that, the more space is occupied by the solid phase, the less is available for the fluid phases, water and air; in other words, the less is the pore space.

444

The Science of Engineering Materials

In the hydration process the specific volume of the solid phase increases. Cement has a density of 3,1 grams/cc, hydrated cement 2.5 grams/cc; and the specific volume of the solid phase increases by about 55 percent, As the hydration progresses the volume of the solids increases and the porosity diminishes, since the total volume is fixed. Throughout almost the entire range of the isotherm, adsorption increases with increasing hydration, because the surface area is the controlling factor. However, close to saturation porosity becomes the controlling factor, and adsorption diminishes with increasing hydration. Powers divided the water in the liquid phase of the hardened paste into two parts: adsorbed water and capillary water, He visualized the former as being under the direct influence of the adsorbing surface, whereas the latter is beyond the range of the surface forces of the solid phase, The adsorption isotherms show no such distinction, and in the foregoing discussion the term adsorbed water included both the adsorbed and the capillary water of Powers. Powers and his co-workers have considerable evidence from permeability, swelling, and freezing experiments to show that in hardened paste the water is held in these two different ways. It is possible that Copelands studies on the thermodynamics of water adsorption will also reveal this. Figure 16.5 shows adsorption isotherms on three pastes of a portland cement, at a definite stage of hydration, with three different wIo/c ratios. The definite stage of hydration means that wn/c, the non-evaporable water per gram of cement, was the same for the three pastes (w./c = 0.15). The rate of hydration of cement increases with increasing wO/c ratio; consequently, the paste with the lowest wo/c ratio (0.316) was the oldest (20 days), and that with the highest wo/c ratio (0.582) was the youngest (7 days). The lower parts of the three isotherms coincide, whereas the upper parts do not, This, again, can be very simply explained in terms of Powers theory of paste structure, The lower parts of the isotherms are determined by the surface area; and it is clear from the figure that the three pastes had the same surface area. This is in agreement with Eq. 16.1; since k is a constant for a given cement and since the three pastes had the same w., therefore, V~ had to be equal for the three pastes, The upper parts of the isotherms are determined by the porosities, which are different for the three pastes. The pastes were obtained by mixing the same weight of cement with different weights of water. After setting, the paste with the highest wO/c ratio had the largest total volume; that with the lowest wo/c ratio had the smallest total volume. Since the volume of the solids was the same for each paste (which follows from

Physics and Chemistry

of Cement

445

the equality of win/c), the volume of the pore space of the paste with the largest wO/c ratio had to be the largest. The adsorption at higher pressures is, therefore, largest for this paste.

[
~

9-6

0,582

0.42

Cement 15007 J wn/c=o.15

? g ~ 0.38 .6 WOIC=0.316L

P tP8

Fig. 16.5. Water-vapor adsorption isotherms on three pastes of a portland cement, having different we/c but the same Wn/c. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute.) On the basis of the above experimentsand hundreds of others Powers and his co-workers have conceived a model of the structure of hardened portland-cement paste. The gel particles are, on the average,

446

The Science oj Engineering Materials

100 or 150A in diameter, and they are roughly spherical in shape. These particles are close-packed in the paste. If the cement gel consisted of equal-sized spheres in hexagonal close packing, the pore volume would be about 26 percent of the total volume. The electron microscopic pictures of Swerdlow and Heckman 18 indicate about a fourfold variation in particle size; in close packing of spheres this would reduce the pore volume to about 20 percent. The water, which is in this pore space, is called the gel water. Expressed in terms of V~, i.e., the weight of water required to cover the entire gel surface with a single adsorbed layer, the gel water would amount to 2 to 3Vm,. This implies that the average size of the gel pores is about 5 times the diameter of a water molecule, or about 15A. In such narrow pores the water is under the immediate influence of the adsorbing surface. The capillary water is visualized as existing in larger pores, between aggregates of gel particles. These larger pores fill up completely only close to saturation pressure. The isotherm in Fig. 16.5 corresponding to wo/c = 0.582, at saturation pressure indicates an evaporable water content equal to about 9vm. Since the gel water is 2 to 3V~, the capillary water is 6 to 7VW. On the other hand, the isotherm corresponding to wO/c = 0.316 has a much smaller pore volume; at saturation pressure there is only 50 percent more capillary water than gel water in this paste. Some pastes are so dense that they contain practically no capillary water. Copeland and Hayes examined old pastes with very low we/c ratios; and for one paste, which hydrated for more than 11 years, they obtained a total evaporable water equal to only about 2.35V~. In this paste all, or almost all, of the water must have been gel water. The entire gel in the hardened paste forms a coherent, porous mass, held together, probably, by chemical bonds. We do not know enough at this stage about the binding of the ultimate particles to each other, but the great strength of hardened paste seems to indicate forces stronger than van der Waals forces between them. In addition,, the gel is a limited-swelling, almost rigid gel. Uptake of water causes only very slight swelling, whereas, it seems that, if the particles were bound by van der Waals forces only, the gel ought to swell much more and, eventually, be dispersed in water. Another school of thought on the shapes of the ultimate particles and the nature of their binding is that of Bernal 20 and his co-workers in England. Bernal believes that the ultimate particles are tiny fibrous crystals; and he thinks that the random crisscrossing and intertwining of the fibers gives hardened paste its rigidity. His ideas are based on

Physics and Chemistry

oj Cement

447

electron photomicrographs obtained by Grudemo for hydrated calcium silicates. Only future work can decide whether Swerdlows or Grudemos pictures represent the shapes of the gel particles correctly or both; however, this has no essential bearing on most of the considerations advanced thus far or on the ones to be advanced later. Chemical composition and structure. So far we have discussed the physical structure of hardened paste; we shall now discuss briefly its chemical composition and structure. From the practical point of view it is the physical structure of hardened portland-cement paste Concrete and mortar are structural mathat is of prime importance. terials, and the structure of the hardened paste in these materials is of decisive importance in determining their structural characteristics, For the builder of buildings, roads, and dams, the important factors are the workability, the strength, the dimensional stability, and the durability of the materials that he uses. It does not matter what chemical compounds the material is composed of, provided that it has the requisite physical propertiesand provided, of course, that the material is not too expensive. For the practical man the chemical constitution of hardened paste is important to the extent that it determines, or influences, or modifies the physical properties of the material. Such practical considerations led, in the first place, to the four modifications of Type I portland cement. Strength considerations led to the development of Type 111 cement, and durability considerations to the other types, as pointed out before. The desired physical characteristics were achieved by alteration of the chemical composition and fineness of normal portland cement, Cements of different chemical composition, naturally, produce pastes of different chemical constitution. The solid phase in hardened portland-cement paste, cured (or hydrated) under water, consists of three identifiable constituents: unhydrated cement, the gel, and calcium hydroxide, The calcium hydroxide comes from the hydration of the two calcium silicates; the gel comes from all the constituents of cement. When the paste is exposed to air, it also contains in the surface layers calcium carbonate, which comes from chemical reaction between atmospheric carbon dioxide and the hydration products of cement. The chemical composition of the solid phase may vary greatly with the type of cement. Starting with the same, wO/c ratio, a one-year-old paste made from a Type III cement tiay contain more than twice as much calcium hydroxide as a Type IV cement paste, and the former may contain only half as much unhydrated cement as the latter. This does not mean, however,

448

The Science oj Engineering Materials

that the two pastes would differ greatly in their important physical propertieson the contrary, the properties are similar to a remarkable extent. Significant differences are found only in the earliest stages of hydrationthe later differences are minor. The most important constituent of the hardened paste is the gel. It is composed of hydrated calcium silicate, aluminate, ferrite, sulfoaluminate, sulfoferriteor, more correctly, it may be composed of some or all of these substances, and possibly others as well. It is possible that the gel is a very fine physical mixture of these substances, and it is also possible that some or all of them lose their identities in the gel. The ultimate gel particles of Powers may be heterogeneous in nature; some may be hydrated calcium silicate particles, others hydrated calcium aluminate particles, and so on. Then again the mixing of the substances may be so thorough as to go down to molecular dimensions; the ultimate gel particles may be copolymers of the hydrates. Whether the hydrated compounds in the gel are copolymers or not, it is remarkable to what a great extent they retain their individual physical and chemical properties. Three examples of this are: (a) Woods, Steinour, and Starke 2 determined the heats of hydration of a number of portland cements having different chemical compositions, and by least-squares analysis they obtained values for the heats of hydration of the four major components, C#3, CW$ C3A, and C4AF. Subsequently, Lerch and Bogue 22determined the heats of hydration of the major components directly. The results of these investigations, as well as those of Verbeck and Foster,z who used the method of Woods, Steinour, and Starke, showed very good agreement. (b) The specific surface areas of hydrated CWSand CYS were directly determined by Grunwald and by Copeland and Hayes, and indirectly by least-squares analysis of data obtained for hydrated cements of different compositions, by Copeland and Hayes, The values were in good agreement. (c) Bogue and Lerch 24 and others determined the water of hydration of the major components of portland cement. Assuming additivity, the author calculated the water of constitution, or w*, for five portland cements, including four different types. Hayes hydrated these cements in a small steel ball mill and obtained complete or almost The experimental and calculated results were in complete hydration. good agreement. The surface area, or V~, the water of hydration, or w., and the heat of hydration, H, are three of the most important basic properties of hardened paste. The fact that these properties can be obtained by

Physics and Chemistry oj Cement

449

adding up the corresponding properties of the major hydrated constituents seems to imply that the gel is a physical mixture of particles of the individual hydrated constituents. This, however, is not necessarily so. Even if the ultimate gel particle is a copolymer of the components, 18

.A I I I I 1/1 / I I 4 /
A
w .
I

I I I
0.12 0.14

2 / 00 - 0.02 0.04 0.06

(S= 120,000 ~ - 3600)

0.08

0.10

0.16

0.18

Fig. 16.6, Relationship between compressive strength and Vm/wO for cements low in C3A. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute.) such additivity is not very surprisingand there is some evidence pointing toward the copolymer hypothesis. For example, Kalousek 25 studied the constitution of hydrated cement by the method of differential thermal analysis. Instead of obtaining separate endothermic peaks for the decomposition of the hydrates of the different components, he obtained only one strong peak, indicating that the gel acted like a single substance. Two other examples will be mentioned later, in the discussion of the hydration process itself, pointing toward the copolymer hypothesis.

450

The Science of Engineering Materials

If we know little about the chemical composition of the gel particles, Are they amorphous we know still less about their basic structure. or crystalline? Even this simple question is difficult to answer. We know somewhat more about the gel particles formed in the hydration of the two calcium silicates, Both C# and CN form the same calcium silicate hydrate, having the composition CS8ZHS (3Ca0 2Si02. 3HzO). When this hydrate is obtained in a well-crystallized form, it has a crystal structure closely resembling that of the natural mineral tobermorite.zo However, when C8S and CZS are hydrated in the form of hardened pastes at room temperature, the X-ray diffraction pattern of the hydrate shows only the two or three strongest lines out of the dozens of lines of tobermorite, and even those are very broad and diffuse. The question therefore arises: are the particles crystalline or amorphous? They are certainly more crystalline than liquids but far less crystalline than metals or known crystalline salts. The same is true of the gel particles of portland-cement paste. Until further information is forthcoming, the calling of the gel particles crystalline or amorphous remains a matter of semanticsand not of physics. However, further information is and will be forthcoming. Vigorous use of such powerful experimental tools of solid-state science as X-ray diffraction, electron diffraction, electron microscopy, differential thermal analysis, thermogravimetric analysis, and others will eventually solve all the problems discussed in this section and many others.
Physical Properties in Relation to Chemistry and Structure

Having discussed the basic structure of hardened paste, it is time now to discuss the relation of such basic properties as surface area, porosity, etc., to such engineering properties as strength, deformation, etc. No spectacular, quantitative, and fundamental correlations have yet been achieved. Our knowledge of the structure of paste is new and far from complete. Our understanding of the basic forces involved in the strength of hardened paste and of the mechanism of strains and stresses is rudimentary. The result is that we now have some empirical correlations and some qualitative explanations. Let us start with strength. In 1919 Duff Abrams G published his great work on the Design of Concrete Mixtures. He considered that his most important result was the relationship that he obtained between the compressive strength of concrete and the mixing water to the cement ratio, He expressed this relationship in the form of the empirical equation

Physics and Chemistry

O) Cement

451

A = A/B

(16.2)

where S is the compressive strength of concrete in psi, x is the waterto-cement ratio expressed in volume units, and A and B are empirical constants. Abrams used one cement and measured the strengths of the 6-by-12-in. concrete cylinders after 28 days of hydration. For this particular case A = 14,000 psi, and B = 7. We may use weight ratios instead of volume ratios, which only changes the numerical value of the constant B. The term x, then, is the same as wO/c in the section on hardened portland-cement paste. We have seen that wO/c determines the porosity of the paste; Eq. 16.2, therefore, gives the relationship between strength and porosity. It is clear from the form of the equation that compressive strength diminishes rapidly with increasing porosity. The strength of concrete depends on the strength of the hardened paste, which in turn depends on the solid matter in the paste; the pores in the paste detract from the strength. Powers and Brownyard 15continued from this point and investigated the relationship between the compressive strength of hardened paste and the solid-matter content of the paste. At the time cement sets, the final volume of the paste is established. As hydration progresses the volume of the solid phase increases, as we have seen in the section on hardened paste. The maximum space available for the increase in the volume of the solid phase is the volume initially occupied by water, which is proportional to WO. As the fraction of the available space occupied by the solid hydration products increases (or as the fractilon of the space occupied by pores decreases), the strength of the paste must increase, The volume of the hydration products is proportional to wn, the nonevaporable water; the fraction of the available space occupied by solid hydration products is, therefore, proportional to wm/wo. Powers and Brownyard found that the dependence of compressive strength on wti/wo was linear, and for a variety of different cements it obeyed the empirical equation ASP = 34,200 (w./wo) 3600 (16.3)

where 8P is the compressive strength of hardened portland-cement paste in psi. They also found that the dependence of SP on V~/wo was linear. Whereas w. is proportional to the quantity of all the hydration products, l~ is proportional to the quantity of the gel in the hydration products. This is because the calcium hydroxide produced in the hydration deposits in larger crystals, which make only a slight contribution to the surface area. The equation for V~ is

452

The Science oj Engineering Materials

s, =

120,000 (V~/wO) 3600

(16.4)

Figure 16.6 shows the dependence of compressive strength on surface area. The relationship is valid for many portland cements, but not for all of them. Cements high in C3A develop lower strengths for a given Vw/wO ratio than the equation indicates. The strengths of these high CSA cements can be brought back to the normal value if more gypsum is added to the clinker, as was shown by Lerch.2T Thus, although strength is primarily dependent on physical structure, the effect of chemical composition cannot be neglected. Later Powerss further developed his ideas on the relation between compressive strength and basic properties of hardened paste. He pointed out that Abrams equation would have a more general validity if x were interpreted as wo/c~ (and not as wO/c), where cn is the quantity of hydrated cement. Since Abrams, in the paper discussed, used only one cement at a definite stage of hydration (28 days), cfiwas proportional to c, and Eq, 16.2 became valid for one particular instance. For a different type of cement at the same age or for the same cement at a different age, clifferent empirical constants A and B must be used~ whereas if % is used instead of c, an equation that is valid for a variety of cements at a variety of ages is obtained. It would be impossible to consider this problem in greater detail. For the present discussion, let it suffice that we now have empirical equations relating compressive strength to such basic properties as porosity and surface area, and that we have a qualitative understanding of these equations. This is a definite advance, if we consider that a few years ago we had no such knowledge. Mechanisms of deterioration. The deterioration of concrete is caused sometimes by chemical and sometimes by physical factors. Examples of the former are reactions with water high in sulfates and reactions of certain aggregates with the alkalies dissolved in the liquid phase of hardened paste. However, even in the alkali-aggregate reaction the primary destructive factor may be physical in nature. Hansen 2 suggested that the reaction of alkali hydroxides with opaline silica in the aggregate produces alkali silicates, which, because of the semipermeable nature of the hardened paste, give rise to osmotic pressures. These osmotic pressures may be responsible for the deterioration of concrete. Most of the trouble with concrete comes from dimensional instability. Volume changes, which set up strains and stresses in the system, may be looked upon as the fundamental physical cause of deterioration,

Physics and Chemistry

O! Cement

453

The volume changes, in turn, are produced by changes in temperature and pressure (the partial pressure of water vapor in the surroundings). Temperature changes may cause volume changes directly (thermal expansion and contraction), or indirectly, by producing phase changes (freezing and thawing). Distress may be caused by the difference between the thermal expansions of aggregate and of hardened paste, or by the expansion of the water on freezing in the capillary pores. Changes in the partial pressure of water produce shrinking and swelling of the gel. The temperature and pressure changes themselves are brought about either by the weather or by chemical and physical processes going on in the system (heats of hydration, adsorption, freezing). The experimental results and the theoretical interpretations of all these phenomena cannot be discussed even briefly. Only two examples will show how the phenomena can be explained in terms of the structure of hardened paste developed before. One of these, shrinking and swelling, seems to be dependent on the gel-water content of the paste; the other, freezing and thawing, on the capillary-water content. Let us start with the shrinking and swelling of hardened paste. When water is removed from the paste, the paste shrinks, and, when water is added, it swells. Aside from an irreversible shrinkage in the course of the first drying of the paste, the phenomenon is reversible. The simplest explanation would be to assume that the pore volume diminishes in proportion to the water removed from the paste. This, however, does not happen. Powers and Brownyard5 showed that the removal of a single molecular layer of adsorbed water, on this basis, would cause a 16 percent decrease in the volume of the paste. In contrast to this, a thin paste slab, with wO/c = 0.5, showed only 0.7 percent linear shrinkage when all the evaporable water was removed, which corresponds to a volume shrinkage of only about 2 percent. Unhydrated cement, calcium hydroxide, and calcium carbonate do not swell on addition of water. The property of swelling resides in the gel, and it is produced by the interaction of gel and water. When water is adsorbed, the surface tension of the gel particles diminishes, and this may be the cause of swellingor, at least, an important factor in it. The mechanism of swelling is not yet clear, but there is ample evidence to show that not all the evaporable water contributes to it. Pickett 80investigated the drying of a concrete prism and showed by mathematical analysis that the results could be explained on the assumption that water was involved in two ways. One of these con-

454

The Science oj Engineering Materials

tributed little or nothing to shrinkage, whereas shrinkage was proportional to the amount of water removed in the other way. On the basis of the quantities involved, Powers and Brownyard identified the former with capillary water, and the latter with gel water. As a matter of fact, it is possible that the major contribution to swelling is made by only a part of the gel water. There is some evidence that gel water itself may be of two types: surface water and interlayer water. Taylor 81 prepared synthetically a calcium silicate hydrate, which was shown to be similar to or identical with the natural mineral tobermorite. Both the synthetic and the natural hydrate are layer crystals, and the distance between the layers changes with increasing or decreasing water content. It was stated above that CSS and CZS, hydrated in the paste form, yield a product that is similar to tobermorite. Bernal 20 and Kalousek believe that portland-cement gel is similar to hydrated C8S and CZS; they believe that the gel has a layer structure and that the distance between the layers changes with changing water content, It was found by Blaine, Hunt, and Tomes3 and by Copeland, Kantro, and others that the surface area of pastes of portland cement and C3S measured by nitrogen adsorption was considerably smaller than the area measured by water adsorption. The discrepancy was roughly twofold; sometimes more, sometimes less. It is clear that water molecules can penetrate to surfaces that cannot be reached by nitrogen molecules. The simplest explanation is that nitrogen is adsorbed only on the outside surfaces of the gel particles, whereas water is not only adsorbed on the outside (surface water) but also penetrates between the layers (interlayer water). Kalousek believes that only the interlayer water is responsible for the swelling of hardened paste. The problcm can be settled only by future research. There is no evidence to date that penetration of water changes the interlayer distance in the cement gel. It is possible that water adsorbed between the layers contributes more to the swelling of the gel than water adsorbed on the outside surface, but the reverse is also possible. So far swelling has been discussed as though it were dependent on physical factors alone. This is far from the whole truth; the chemical composition of the gel plays an important role in swelling. Gonnerman w and Woods, Starke, and Steinour 5 investigated the effect of the chemical composition of various cements on the shrinkage of mortars made with these cements. They found that the shrinkage of the mortars correlated well with the C3A content of the cement and not with the CSS and CZS content. It is clear from this that the swelling

Physics and Chemistry of Cement

455

of hardened portland-cement paste cannot be explained by simply postulating that it is analogous to the swelling of hydrated calcium silicates. Atmospheric carbon dioxide also plays an important role in the shrinking of hardened paste; however, the mechanism of this action has not been elucidated as yet. The last topic for discussion about hardened paste is freezing and thawing. Just looking at some concrete roads after years of hard winter freezes makes one realize the importance of the destructive effect of the freezing and thawing of water in concrete, Yet, as will be shown shortly, this problem is solvedin principle completely, in theory and practice partly-and in the coming years freezing and thawing will cease to be an important problem in the use of concrete. The solution of the problem came by accident. The pore space of hardened paste is, of course, filled with water and air. If we artificially entrain more air in the paste, in the form of air bubbles in the proper sizes and the proper distribution, the strains and stresses set up by freezing and thawing are greatly diminished, and concrete becomes vastly more durable. The most widely used air-entraining agents are Vinsol resin and Darex. We can explain the mechanism of air entrainment. Powers and Brownyard 5 began to tackle the problem of the freezing of water in hardened paste, They noticed that some of the water in the pore space froze and some did not; and they found that in the temperature range from 6 to 15 C, the freezable water roughly corresponded to the capillary water. They concluded that the water did not freeze in the gel-pore spaceor at least it did not freeze down to 78 C. This is a reasonable conclusion in the light of the theories of crystal formation. We believe that crystals are able to grow in a melt or in solution only if the nuclei exceed a certain critical size. We have seen how exceedingly narrow the pore space is between the gel particles; it is possible that the critical size of an ice nucleus in a gel pore is greater than the available dimensions, It is also true, quite aside from critical size, that an adsorbed film freezes at a much lower temperature than the three-dimensional liquid. It seems clear, therefore, that at winter temperatureseven in a very hard winterwater freezes only in the capillary pores. On the basis of this fact, Powers developed a hypothesis to explain frost action and also the mechanism of the effect of air entrainment. Figure 16.7, taken from Powers and Helmuth,7 gives a diagrammatic representation of the structure of hardened paste, discussed before. The gel particles are shown as black spots, and the capillary pores

456

The Science of Engineering

Materials

The curved boundary at the left represents a part of the wall of an air bubble, and it shows that the dimensions of such entrained air voids are gigantic compared with the dimensions of capillary pores. In a saturated paste both the gel pores and capillary pores are ful~ of water. When the temperature falls to a point at which freezing can begin, ice appears, first in the largest capillary pores. Let us say

as empty spaces,

Gel particle>

Gelpore
Boundary

of part of
an air void

Fig. 16.7. Diagrammatic representation of gel structure with air void. (From Powera and Helmuth. By courtesy of the authors and the Highway Research Board.) that (7I is such a pore. When water freezes, it expands, since the volume of ice is 9 percent greater than that of water, This expansion will cause either dilation of the pore or expulsion of some water from it. Cement paste is somewhat permeable to water, though the permeability is extremely low. If there is an air void not too far from Cl, the excess water may escapevia the gel poresinto the air void. The growing body of ice acts as a pump that forces water through the paste toward the void boundary, thus generating a hydraulic pressure in the system. The most important factors controlling this pressure are: (1) the coefficient of permeability of the paste; (2) the distance from G1to the void boundary; and (3) the rate of freezing. If we now look at capillary pore Czin Fig. 16,7, it is clear that water can escape from it more easily than from Cl, since it is much closer to the air-void boundary. The hydraulic pressure generated by

Physics and Chemistry

of Cement

457

freezing in Cz must be less than that generated in Cl. The farther away is the escape boundary, the greater the pressure. If the boundary is too far, the pressure may stress the surrounding gel beyond its elastic limit, and thus may cause permanent damage, Experiments proved beyond question that Powers theory is correct hydraulic pressure plays an important role in frost action, However, it is not the whole story. Collins8 advanced a different theory to explain the destruction of concrete by frost, At first, the two theories were regarded as alternatives, and there were certain experiments that could be explained only on the basis of Powers theory. Later, however, the complex experimental results obtained by Powers and Helmuth convinced them that the hydraulic pressure theory alone could not account for all the results, but that the two theories together could account for all the observations. Hydraulic pressure seems to be the most important factor in producing (salt scaling, i.e., the scaling of concrete pavements produced by salts used to thaw the ice and snow on pavements in winter. Verbeck, Klieger, and Gramlich have found that the presence of a salt significantly decreases the coefficient of permeability of hardened paste, This should result in greater hydraulic pressures and more deterioration than would be produced by freezing in the absence of the salt. This chapter is only a brief outline of some of the things we know and believe about the physical and chemical structure of hardened paste. Only a few equations have been introducedbut do not gather from this that the information is almost entirely qualitative. A great mass of quantitative information exists, of which only a small fraction is given. Those who like equations should consult some of the referencesPCA Bulletin %2 alone contains hundreds of equations. Some of the subjects that may be of great interest to engineers and scientists have been scarcely touched upon. Such, for example, is the permeability of hardened paste to water. Hardened portlandcement paste is less permeable to water than some stones and rocks. For example, a sample of granite was found to be more permeable than paste, which is so much the more surprising since the former had only 1 percent pore space, whereas the latter had about one-third pore space. An interesting experimental result obtained from permeability measurements is the confirmation of the surface area of hardened paste measured by water adsorption. Carman9 developed a method to determine surface areas of macroscopic particles by permeability experiments. Powers, Copeland, and Hayes, who were the first to use the method for such a fine, colloidal system, obtained permeability

458

The Science oj Engineering Materials

surface areas for pastes that checked the adsorption areas reasonably well. The model of the structure of hardened paste, developed by Powers and his co-workers, is the first quantitative approximation model. Further research will modify and enlarge this model, and eventually a second approximation model will be built. This future model will give more information about the size and shape distribution of the gel particles, the pore size distribution, and the distribution of the surface areas within the various pores. It will also tell more about the chemical composition and crystal structure of the gel particles. From such a model more fundamental correlations between paste structure and engineering properties will be drawn. For example, on the subject of dimensional instability, so far all investigations examined overall volume changes, but the deteriorating effects of volume changes do not necessarily correlate with overall changes, The rate It is also clear that large localized of change may be of importance. volume changes may produce large strains and stresses, even if the overall volume change is small. Thus, a more detailed knowledge of the structure of hardened paste would help in understanding better the nature of destructive deformations. The second approximation model on which we are working is still only a physiochemical model. Eventually a third one will follow, probably based on solid-state science. The time will come when solidstate science will be able to tackle even as complex systems as are being discussed now. Then a really fundamental understanding of hardened portland-cement paste will have been achieved.
The Hydration of Portland Cement

Some aspects of the hydration of cement will now be discussed. The investigation of the hydration reaction, like that of any other chemical reaction, consists of two phases: (a) Investigation of the kinetics of the hydration process, i.e., a study of the course of the chemical reaction itself, while it is in progress. (b) Investigation of the thermodynamics of the reaction, i.e., a study of the system after equilibrium has been reached and no further reaction takes place. There is very little information available on (b), simply because the reaction is very slow, and real thermodynamic equilibrium is probably not reached by the hydrating system even in a decade,

Physics and Chemistry

O) Cement

459

Therefore, the discussion will be restricted to the rate of the hydration process. As a zeroth approximation to the true state of affairs, we may assume that each of the constituents of portland cement, including major and minor components, hydrates independently of the others. In this event, the overall rate of the reaction should be the sum of the rates of hydration of the different components. As a matter of fact, this assumption is known to be false for at least two of the components, CSA and gypsum. Gypsum diminishes strongly the rate of hydration of CSA, as was shown by Lerch.27 By itself, C3A hydrates very rapidly, and in the absence of gypsum it causes a rapid setting of cement (a flash-set). Gypsum is added to the clinker to slow down the hydration of CSA, which it does by reacting with C3A and water to form a hydrated calcium sulfoaluminate. Even though the assumption of independent rates cannot be completely true, it is legitimate to inquire whether it comes close to the truth. We are in a position to give at least an approximate answer, since we have information about the rates of hydration of the major components. The most important work in this field is that of Bogue and Lerch.24 They investigated the rates of hydration of the major constituents individually and mixed with gypsum, and they also investigated the rates of hydration of some mixtures of the major components. The rates were followed to the age of 2 years by determining the water of constitution, or fixed water,l w~. I?igure 16.8 represents the fractions of the four major components hydrated from 1 day to 100 days.40 Indirect determinations of the rates of hydration of the individual components are found in the papers of Woods, Steinour, and Starke 21 and of Verbeck and Foster.2S These investigators determined the heats of hydration of portland cements of different compositions at different ages and, by least-squares analysis, evaluated the heats of hydration of the major constituents at clifferent ages. Assuming that the heats were proportional to the amounts hydrated, there seemed to be at least a rough additivity of the rates of hydration of the major components. This appeared to be the conclusion from the results of Bogue and Lerch as well. Copeland and Bragg determined the rate of hydration of a Type I portland cement and compared it with the rate calculated from the data of Bogue and Lerch, of Woods, Steinour, and Starke, and of Verbeck and Foster, on the assumption that the rates of the individual components were additive. They concluded that the agreement was

460

The Science oj Engineering Materials

poor. This is not surprising, partly because of the already-mentioned effect of gypmm on the hydration of CSA, partly because the results of Bogue and Lerch for mixtures showed that some of the components had some effect on the hydration of others. The hydration of the calcium silicates seemed to be somewhat accelerated by C3A and retarded by C4AF. Let us examine now a different hypothesis for the explanation of the rate of hydration of portland cement. The starting point of this l.o , 0.8, y-cJ ~ -g 0.6 .c c 0 . 0.4 5
P IL

C4AF C3A

0 -3

C3S

0.2~

o~

c#3

10
Time (days)

100

Fig. 16.8. Rates of hydration of pure compounds. (From Copeland and Bragg.) hypothesis is the relationship land-cement paste. stated in the section on hardened portv. = kwn (16.1)

Figure 16.9, taken from Powers and Brownyard,s illustrates this relationship. The points represent the V~ and w. values of different pastes made from the same cemenl+the pastes had different wO/c ratios and different ages. For another cement a different straight line is obtained, also passing through the origin, but the slope of the line k is different. V~, representing the surface area, is proportional to the quantity of colloidal material in hardened paste; Wn, representing the water of hydration, is proportional to the quantity of all hydration products, colloidal and non-colloidal, The constancy of V~/wn implies that the same kind of hydration products are formed at all stages of hydration. This was pointed out by Powers and Brownyard. Copeland and Bragg 40 went a step farther by pointing out that the constancy of

Physics and Chemistry of Cement

461

V~/wn implies that the fractional rate of hydration of all components of a given cement is the same, In other words, when a cement is 50 percent hydrated, the CWSin it is 50 percent hydrated, the CYS is 50 percent hydrated, the CSA is 50 percent hydrated, and so on. The constancy of V~,/w,, for a given cement is a surprising experimental fact. If we look again at Fig. 16.8, giving the results of Bogue
007

0.06

0.05

0.04 y,, 0.03

0.02

0.01

0,

Wn

Fig. 16.9. Vmvs. Ww relationship for pastes of the same port,land cement. (From Powers and Brownyard. By courtesy of the authors and the American Concrete Institute.) and Lerch 24 on the individual rates of hydration of the four major components, we see how different the rate of hydration of Cfi is from that of C3A. It is not difficult to imagine that in the presence of each other the rate of hydration of C2S is accelerated and that of C3A is retarded, but to find that the fractional rate of hydration of CZS and C3A is the same, is, to say the Icast, surprising. And yet this is what the constancy of k implies. We can see this better if we consider a simple cement, consisting of C3S and C2S only. The relation V~/wn is proportional to the gel fraction of the hydration products. The gel fraction is quite different for C3S and C2S. In the hydration of C3S and C2S the same calcium silicate hydrate, or gel, is produced, but the former produces one more mole of Ca (OH) ~ per

462

The Science of Engineering Materials

mole of reactant than the latter. Some experiments by Kantro, Pressler, and the author have confirmed previous indications that Ca (OH) z makes very little contribution to the surface area. From this it follows that V~/wn for hydrated C&l should exceed that for hydrated CS8 by about 50 percent. Now let us assume that in our hypothetical cement C3S hydrates at a higher fractional rate than CZS; for example, when our cement is 20 percent hydrated, the CZAin it is 5 percent hydrated, and the CSS in it is 25 percent hydrated. We would then find, on the basis of the foregoing, that k would increase with age, since with increasing age the ratio of hydrated CZS to hydrated C3Swould increase. However, we do not find such an increase; k remains constant, which implies that the fractional rate of hydration of the components is approximately the same. Two different hypotheses of cement hydration have been discussed, the hypothesis of independent hydration of the individual components and the hypothesis of equal fractional rates, and having preferred the latter, we may inquire how wrong is the former? Could we obtain, at least roughly, the observed dependence of V~ on W. by assuming independent hydration of the constituents? Figure 16.10, taken from Copeland and Bragg, gives the answer to this question. l~ and W. were determined for more than a hundred different pastes prepared from the same cement, and the least-squares line was drawn as shown in the figure. (Only a few of the experimental points are shown.) The two lines that flank the least-squares line represent the 95 percent confidence limits calculated from the data. This means the following: if we have many series of pastes of this cement, with a definite value of wn, and calculate the mean value of l~ for each of these series, we expect that 95 percent of the mean values will fall within those limits. Using certain assumptions, and assuming independent rates of hydration of the individual components, Copeland and Bragg calculated the V~ vs. W. relationships from the data of Bogue and Lerch,24 of Woods, Steinour, and Starke,zl and of Verbeck and Foster.23 These curves are shown in Fig, 16.10, and all three fall outside the 95 percent confidence limits. Copeland and Bragg concluded, therefore, . . . that the components in a paste of portland cement do not hydrate at rates characteristic of their pure states, Some of the assumptions on which these calculations have been based are not unimpeachable, and Fig. 16.10 does not definitely exclude at least a rough additivity of the rates of the components. The hypothesis of equal fractional rates was so surprising as to be difficult to believe without independent corroborative evidence. Cope-

Physics and Chemistry

of Cement

463

land and Bragg 40pointed out that such independent evidence in data on heats of hydration. The heats of hydration of the four major components differ the lowest is that of CZS, about 60 cal/gram, the highest is that In spite of this, if all the components of about 200 cal/gram.

existed widely; of C3A, a given

0.07

0.06

O.OE

0.04 2 >s 0,03

0,02

0,01 0
cum /c

Fig. 16.10. Experimental and calculated Vm/C vs. wJc relationship for pastes of a portland cement. (From Copeland and Bragg,) cement hydrate at the same fractional rate, not only should V~ be proportional to wn, but Hn, the heat of hydration of the cement, should also be proportional to wn, H. = ]CfW. (16.5) .

If the nature of the hydration products in the earlier stages of hydration is the same as in the later stages, the heat evolved per unit weight of hydrated material, which is proportional to Hn/wm, should be constant at all stages.

464

The Science oj Engineering Materials

Verbeck and Foster 29demonstrated proportionality between H. and w.. The author has analyzed the data of Verbeck and Foster by plotting H. vs. w. for each of the 24 portland cements that they had investigated, the hydration periods ranging from 7 days to 61/2 years. He found that Eq. 16.5 held approximately for all of their Type I and
I
120

q Type III and 111A

100

(4 cements) OType I and IA (12cements)

80 M z s ~

60-

40

20

Fig. 16.11. Relationship between H/c and wn/c for hydrated Type I and Type III portland cements, (Based on the data of Verbeck and Foster.)

Type 111 cements and almost exactly for some of them; and k was almost the same for all 12 Type I and 4 Type III cements. The composite curve for these 16 portland cements is shown in Fig. 16,11, You can see how accurately Eq, 16.5 is obeyed for the averages of the experimental data. The figure also shows the Hn vs. W. curve for CWS. The two curves are close to each other, the cements having somewhat higher heats of hydration for a given W% than C&. A similar analysis of Verbecks data on Type II, IV, and V portland cements proves that the equal fractional rate hypothesis does not apply to them. These cements are higher in CXS content than the other two types. In the earlier stages of hydration the slope of the H. vs. W.

Physics and Chemistry oj Cement

465

curve for Type II, IV, and V cements approaches that of C3S; in the later stages it approaches that of C2S. In the hydration of Type I and III portland cements, which are relatively lower in CZS, the hydration of the other components carries along the hydration of (72S, so that approximately equal fractional rates are obtained; whereas with cements high in Czfithis happens only in the earlier stages of hydration, and eventually some C2S remains to hydrate by itself. The author is now almost fully convinced that the equal fractional rate hypothesis of Copeland and Bragg is valid for Type I and Type III portland cements. It may or may not be valid during the first day of hydration, but it seems to be valid from 1 day to more than 11 years. None of the Portland Cement Association research chemists is as yet one hundred percent convinced of the correctness of the hypothesis. Copeland and Bragg are working on another and more direct test of the hypothesis: they are determining, by means of an X-ray spectrometer, the quantities of unhydrated compounds in pastes of various ages made from the same cement. The heats of hydration having been discussed in the previous section and elsewhere, it is important to clarify now the meaning of the term as used throughout this chapter. Powers and Brownyard pointed out that the heat evolved in the hydration of portland cement is a composite quantity, consisting of two parts: the heat evolved in the hydration reaction, which we may call the chemical heat of hydration, and the heat evolved in the adsorption of water on the gel produced in the hydration. All the heat data presented in this chapter were overall heats of hydration. Powers and Brownyard determined the heat of adsorption of water on two different cement pastes, and they came to the conclusion that about one-fourth of the overall heat of hydration was heat of adsorption. J3runauer, Hayes, and Hass M determined the chemical heats of hydration of C,IS and CJ3, and on the basis of comparison with the data of Verbeck and Foster 23 concluded that about one-fourth of the overall heat of hydration of C8S was heat of aclsorption, but more than half of the overall heat of hydration of CZS was heat of adsorption. Further work on the chemical heats of hydration of C3S and Czt$ has convinced the author that the overall heat of hydration was even more [composite than had been thought before. The gel formed in the hydration has a tremendously large surface area, and the surface energy of the gel represents a sizable fraction of the energy changes occurring in the system. The chemical heat of hydration of CZS,for example, may vary by a factor of 2, depending upon the specific surface

466

The Science of Engineering Materials

area of the hydrate formed. The heat of adsorption of calcium hydroxide on the gel may also be a factor, although to date we have no experimental evidence to support this. Let us return now for a moment to the chemical nature of hardened paste. In the previous section, the question was raised whether each of the ultimate gel particles was a single chemical compound only, or whether it was a copolymer of many compounds. Three lines of investigations were mentioned pointingthough vaguelyin the direction of the gel being a physical mixture of particles of the individual compounds, and one set of investigations was mentioned pointing in the opposite direction. The constancy of V~/wm and of H./w., discussed in this section, point in the direction of the copolymer structure. The evidence for the copolymer structure is somewhat stronger than the evidence for the mixture. Future research will soon make a definite decision. Swerdlow and Heckman at the Bureau of Standards are trying to focus their electron beams on individual gel particles and to obtain electron diffraction patterns of the particles, The answer will soon be known. The basic chemistry and physics of concrete and motiar have not The three-component system is, naturally, more been discussed. complex than the two-component system; consequently, our understanding of it is less complete. Much of the behavior of concrete and mortar can be interpreted at least qualitatively, in terms of Powers model of the structure of hardened paste. This point was illustrated earlier. For a more detailed understanding we must consider the effects of the aggregates. Much important and interesting work has been done in this field, but a summarizing of this work must wait for someone who is better acquainted with it than the author. In conclusion, as no doubt other contributors have pointed out in their chapters, the science of cements is by no means complete. But already we have seen how its pursuit has been translated into important advances in highway technology and in concrete applications. One can only hope that the further application of the methods of chemistry and physics will accelerate further understanding of the material and practical useful developments. I should like to express my appreciation to Mr. T. C. Powers, Dr. H. H. Steinour, Mr. G. J. Verbeck, Mr. Douglas McHenry, and Dr. L. E. Copeland for reading this chapter in manuscript and making valuable suggestions.

Physics and Chemistry

O) Cement

467

BIBLIOGRAPHY
F. M. Lea and C, H. Desch, The Chemistry oj Cement and Concrete, second edition, St. Martins Press, New York, 1956. R. H. Bogue, The ChemistrV of Portland Cement, second edition, Reinhold Publishing Corporation, New York, 1955. Hans Kuhl, Zement Chemie, 3 volumes, Verlag Technik GMBH, Berlin, 19511952. Symposium on the Chemistr~ of Cements, Stockholm, 1938, Ingeni6rsvetenskapsakademien, Stockholm, 1939. Proceedings oj the Third International S~mposium on the Chemistry of Cement, London 1962, Cement and Concrete Association, London, 1954.

REFERENCES 1. R. H. Bogue, The Chemistry of Portland Cement, second edition, Reinhold Publishing Corporation, New York, 1955. 2. M. A. Swayze, Am. J. Sci. 244, 1; 65 (1946). 3. R. H. Bogue, Znd. Eng. Chem., Anal. Ed. 1, 192 (1929). 4. L. A. Dahl, Rock F%oducts 60, No. 1, 109; No. 2, 107; No. 3, 92; No. 4, 122 (1947) , 6. J. W. Jeffery, The Tricalcium Silicate Phase: R-oceedinos oj the Third International Symposium on the Chemistry OJCement, London, 196.2, Cement and Concrete Association, London, 1954, p. 30. 6. R. W. Nurse, The Dicalcium Silicate Phase/ Proceedings oj the Third International Symposium on the Chemistru of Cement, London, 1968, Cement and Concrete Association, London, 1954, p. 56. 7. Fred Ordway, Tricalcium Aluminate: Proceedings oj the Third International Symposium on the Chemietw oj Cement, London, 196%,Cement and Concrete Association, London, 1954,p. 91. 77 (1938). 8. W, Lerchj J. Research I?atl. Bur. Standards 2?0, 9. For example, Bulletins .2,3, and 4 of the Research Department of the Portland Cement Association, (Bulletin ,2 by T. C. Powers, with an appendix by L. A. Dahl, Bulletins .9 and 4 by H. H. Steinour.) 10. H. Le Chfitelier, Compt. rend. 94, 13 (1881). 11. W. Michaelis, Tonind. Ztg, 16, 105 (1892). 12. L. Jesser, Zement lf3,741 (1927); 18,161 (1929). 13. S. Giertz-Hedstrom, Zement .20,672 (1931). 14. E. Freyseinet, Science et ind., Jan., 1933. 15. T. C. Powers and T. L. Brownyard, J. Am. Concrete Inst., Oct., 1946April, 1947, Proceedings .@, 1947, The nine papers were issued also as Bulletin 22 of the Research Department of the Portland Cement Association. 16. S, Brunauer, P. H. Emmett, and E. Teller, J. Am. Chem. SOG 60, 309 (1938). 17. S. Brunauer, L. S. Deming, W. E. Deming, and E. Teller, J. Am. Chem. Sot. 6,?,1723 (1940). 18. M. Swerdlow, H. F. McMurdie, and 1. A. Heckman, Proceedings oj the International Conference on Electron Microscop~, London, 1954.

468

The. Science oj Engineering Materials

19. L. E. Copeland and J. C. Hayes, ASTM Bulletin 194, 70 (December, 1953). 20. J. D. Bernal, The Structures of Cement Hydration Compounds: Proceedings of the Third International Symposium on the Chemistr~ oj Cement, London, IW%,Cement and Concrete Association, London, 1954, p. 216. 21, H. Woods, H, H, Steinour, and H. R. Starke, Znd. Eng. Chem. 24, 1207 (1932); Eng. News-Record 110,431 (1933). 22. W. Lerch and R. H. Bogue, J. Research Natl. Eur. Standards 1.2,645 (1934). 23. G, J. Verbeck and C. W. Foster, Proc. Am. Sot. Testing Materials 60, 1235 (1950). 24, R. H. Bogue and W. Len+ lnd. Ilng. Chem. 26, S37 (1934). 25. G, L. Kalousek, C. W. Davis, Jr., and W. E. Schmertz, Proc. Am. Concrete

Inst. 46,693 (1949); .@,77 (1951). 26. Duff A. Abrams, Design of Concrete Mixtures, Bulletin 1 of the Structural
Materials Research Laboratory of Lewis Institute, April, 1919. 27. W, Lerch, Proc. Am. Sot. Testing Materials 16, 1251 (1946). 28. T. C. Powers, ASTM Bulletin 168, May, 1949. 29, W. C, Hansen, Proc. Am. Concrete Znst. 40, 213 (1944). 30. G. Pickett, Research Reports of the Portland Cement Association Research Laboratory, Appendix 3, July, 1942, unpublished. 31. H. F. W. Taylor, J. Chem. Sot. 1960, 3682. 32. G. L. Kalousek, R. J. OHeir, K. L. Ziems, and E. L. Saxer, J. Am. Concrete

Inst. .26 (3), 225 (1953). 33. R. L. Blaine, C. M. Hunt, and L. A. Tomes, P~oc. O) the Highwa~ Research Board, Jan., 1953, p, 298; L. A. Tomes and C. M. Hunt, Natl. Bur. Standards Report ,2966 (Dec., 1953). 34. H. l?. Gonnerman, Proc. Am. Soc. Testing Materials S4, II, 244 (1934). 35. H, H. Woods, H. R, Starke, and H. H. Steinour, Rock Products W, 24 (1933). 36. T. C. Powersl Proc. Am. ConcreteInst. 41,245 (1945). 37. T. C. Powers and R. A. Helmuth, Proc. Highwag Research Board W, 285 (1953). 38. A. R. Collins, J. Inst. Civil Engrs. 23, 29 (1944).
39. P. C. Carman in Symposium on New Methods of Particle Size Distribution in the Subsieve Range: A.S.T.M. Publication, 1941, p. 24. 40. L. E. Copeland and R. H. Bragg, Determination of Ca(OH)2 in Hardened Pastes with the X-Ray Spectrometerfl May 14, 1953 (Portland Cement Association Report, unpublished). 41, S. Brunauer, J. C. Hayes, and W. E. Hass, J. Phvs. Chem. 68, 279 (1954).

Das könnte Ihnen auch gefallen