18
Linking Fly Ash Composition to Performance in Cementitious Systems Tandré Oey 1 , Cynthia Huang 1 , Ryan Worley 1 , Sarah Ho 1 , Jason Timmons 1 , Ka Lam Cheung 1 , Aditya Kumar 1 , Mathieu Bauchy 1 , and Gaurav Sant 1 1 Department of Civil and Environmental Engineering, University of California, Los Angeles, CA; 2 National Institute of Standards and Technology, Gaithersburg, MD KEYWORDS: Fly Ash, Characterization, Reactivity, Cement ABSTRACT Fly ash represents a critical, yet underutilized resource for use as a supplementary cementing material (SCM) for replacement of ordinary portland cement (OPC) to produce low CO 2 footprint concretes. However, compositional variability and lack of rigorous classification methods currently limits its use, due to unpredictable performance of fly ash in cementitious mixtures. This research evaluates a wide range of fly ashes sourced based on compositional variation as well as abundance. Ashes are exhaustively characterized, and compositions of both crystalline and glassy phases are determined through a combination of XRF, XRD, and SEM-EDS techniques. The compositions are then distilled to a parameter, the network ratio, which is derived from expected network behavior of the alkali/alkaline earth aluminosilicate glass component, which makes up the majority of each ash. This parameter is validated by multiple methods of its calculation, and its correlation to known composition-dependent glass properties such as glass transition temperature and amorphous XRD peak position is demonstrated. Further, isothermal calorimetry, set measurement, and compressive strength determinations are used to evaluate the impact of ash composition on reaction kinetics, and both early age and mature performance of fly ash-cement pastes. A clear link is drawn showing that glasses with high network ratio (i.e. less stable structure) are more reactive, as illustrated by the associated impact on kinetics and property development. 1.0 INTRODUCTION Environmental concerns are increasingly favoring the replacement of ordinary portland cement by supplementary cementing materials (SCMs) in concrete mixtures. 1,2 Fly ash, a waste byproduct from coal burning power plants, is one of the most promising SCMs, as it is relatively abundant, can be used up to replacement levels exceeding 50%, 3 and has potential as a geopolymer precursor which utilizes 100% fly ash (i.e. an alternative binder to cement). 4 Unfortunately, the use of fly ash as an SCM at particularly at high replacement levels can result in retardation of cement hydration, delayed set time, and

Paper-WOCAConference-TJO-03132015-Final

Embed Size (px)

Citation preview

Page 1: Paper-WOCAConference-TJO-03132015-Final

Linking Fly Ash Composition to Performance in Cementitious Systems

Tandré Oey1, Cynthia Huang1, Ryan Worley1, Sarah Ho1, Jason Timmons1, Ka Lam Cheung 1, Aditya Kumar1, Mathieu Bauchy1, and Gaurav Sant1 1Department of Civil and Environmental Engineering, University of California, Los Angeles, CA; 2National Institute of Standards and Technology, Gaithersburg, MD KEYWORDS: Fly Ash, Characterization, Reactivity, Cement ABSTRACT Fly ash represents a critical, yet underutilized resource for use as a supplementary cementing material (SCM) for replacement of ordinary portland cement (OPC) to produce low CO2 footprint concretes. However, compositional variability and lack of rigorous classification methods currently limits its use, due to unpredictable performance of fly ash in cementitious mixtures. This research evaluates a wide range of fly ashes sourced based on compositional variation as well as abundance. Ashes are exhaustively characterized, and compositions of both crystalline and glassy phases are determined through a combination of XRF, XRD, and SEM-EDS techniques. The compositions are then distilled to a parameter, the network ratio, which is derived from expected network behavior of the alkali/alkaline earth aluminosilicate glass component, which makes up the majority of each ash. This parameter is validated by multiple methods of its calculation, and its correlation to known composition-dependent glass properties such as glass transition temperature and amorphous XRD peak position is demonstrated. Further, isothermal calorimetry, set measurement, and compressive strength determinations are used to evaluate the impact of ash composition on reaction kinetics, and both early age and mature performance of fly ash-cement pastes. A clear link is drawn showing that glasses with high network ratio (i.e. less stable structure) are more reactive, as illustrated by the associated impact on kinetics and property development. 1.0 INTRODUCTION Environmental concerns are increasingly favoring the replacement of ordinary portland cement by supplementary cementing materials (SCMs) in concrete mixtures.1,2 Fly ash, a waste byproduct from coal burning power plants, is one of the most promising SCMs, as it is relatively abundant, can be used up to replacement levels exceeding 50%,3 and has potential as a geopolymer precursor which utilizes 100% fly ash (i.e. an alternative binder to cement).4 Unfortunately, the use of fly ash as an SCM at particularly at high replacement levels can result in retardation of cement hydration, delayed set time, and

Page 2: Paper-WOCAConference-TJO-03132015-Final

reduced early age strength development.5,6 As such, current efforts to utilize fly ash effectively must include careful consideration during mixture proportioning,7 which could be expedited or avoided by better initial characterization. To this end, recent studies have recognized the need for improved characterization of both the crystalline and amorphous phases of fly ash, which is quite heterogeneous and prone to variability due to its origin as a waste byproduct of coal burning power plants.8,9 Current standards,10 classifying ashes as Class F or Class C based on bulk oxide composition, fall far short of providing a descriptive characterization of ash to allow for prediction of reactivity and performance as an SCM. While a great many studies have investigated the performance of fly ash in portland cement concrete,11–16 fewer have investigated the link between ash composition, reactivity, and performance,17–19 or engaged in detailed analysis of the amorphous phases.8,9,18,20 The aim of this study is to demonstrate the merit of more comprehensive characterization methods which are applicable across a wide range of ashes, and to bridge the gap between composition, structure, and reactivity. To accomplish this, each of seven fly ashes is characterized with regard to both crystalline and amorphous phase compositions. Reactivity and property development of cement-fly ash blends is evaluated, and links between performance and compositional variations are drawn. The outcomes of this research shed light on the link between glass structure and properties for fly ash, and provide a basis for its widespread and informed as an SCM. 2.0 MATERIALS AND EXPERIMENTAL METHODS An ASTM C150 compliant Type I/II ordinary portland cement (OPC) was used in this study.10 Mineralogical composition of the OPC was estimated using quantitative X-ray diffraction (QXRD) and Rietveld refinement: 56.5% MIII-Ca3SiO5, 16.0% β-Ca2SiO4, 6.3% Ca3Al2O6, 11.4% Ca4Al2Fe2O10, 1.1% CaSO4∙ 2H2O, 0.5% CaSO4∙ 0.5H2O, 1.2% CaSO4, 1.2% Ca(OH)2, 0.5% CaO and 4.6% CaCO3. The seven fly ashes used are commercially available products of Headwaters Incorporated (South Jordan, Utah)1, and were chosen to encompass representative, yet abundant U.S. fly ashes. This set includes three Class C ashes, and four Class F ashes, as described in ASTM C618.10 Particle size distributions (PSD) of the cement and the ashes were measured using a Beckman Coulter light-scattering analyzer (LS13-320, Figure 1) using isopropanol and sonication for dispersing the powders to primary particles. The uncertainty in the light-scattering analysis was determined to be approximately 6% based on six replicate samples, assuming the density of the cement to be 3150 kg/m3.

1 Certain commercial materials and equipment are identified to adequately specify experimental procedures. In no case does such

identification imply recommendation or endorsement by the University of California, Los Angeles or NIST, nor does it imply that the items identified are necessarily the best available for the purpose.

Page 3: Paper-WOCAConference-TJO-03132015-Final

Figure 1: Particle size distributions for the cement and seven fly ashes used in this study. The uncertainty in measured particle size distribution is approximately 6%. 2.1 Characterization of Fly Ashes Each ash was exhaustively characterized with regard to both crystalline and amorphous phase compositions, and other material properties as follows. The bulk oxide composition of each fly ash was determined using X-ray fluorescence (XRF, Table 2). Nitrogen BET surface area measurement was performed using a Micrometrics ASAP 2020 BET apparatus. Heat capacities were measured using a Hot Disk Thermal Constants Analyzer via the transient plane source method. Ash densities were measured using a Micrometrics Accupyc II 1340 Gas Pycnometer, with helium as the purge gas. Loss on ignition and glass transition temperature of the amorphous phase fraction were determined by thermal analysis (TGA/DTG/DTA) using a TA Instruments SDT Q600 Thermal Analyzer, and a heating rate of 10°C/min. Crystalline phase composition was determined by QXRD on a Bruker D8 Advance diffractometer using Rietveld refinement via the commercially available TOPAS program (Table 1). Powders were scanned from 5-70 degrees two theta with a six second step time using a zinc oxide internal standard. Samples were prepared by grinding with ethanol in a McCrone Micronising Mill for homogenization, followed by vacuum filtration and drying. Powders were manually ground with ~20% ZnO internal standard in ethanol, and re-dried prior to mounting. Care was taken during mounting to avoid preferred orientation of crystalline phases. Amorphous phase compositions were measured from carbon coated polished sections using an FEI Quanta 600 scanning electron microscope (SEM) in high vacuum mode. In addition to backscatter images, element maps and point compositions were obtained using X-ray dispersive spectroscopy (EDS). Image analysis was performed in ImageJ and Multispec, with compositional clustering methods modeled after those used by Chancey et al.8,9 GCD kit software was used as an additional checkpoint for both manual and automated clustering analyses.21

Page 4: Paper-WOCAConference-TJO-03132015-Final

2.2 Evaluation of Ash Reactivity and Performance Cementitious paste mixtures were prepared with deionized water at a water to solids ratio of 0.45 by mass, using a planetary mixer per ASTM C305.10 Fly ash was added by replacement of cement at levels of 10, 20, and 50% by mass. Influence of fly ash inclusion on reaction rate was measured using isothermal calorimetry at 25°C on a Tam Air calorimeter (TA Instruments). Compressive strength of 5 cm cube samples cured at 25°C in saturated lime solution was measured per ASTM C109 at 1, 3, and 7 days.10 Initial and final set time of pastes was determined by vicat needle testing per ASTM C191.10 Reported values for both set and strength are the average of three samples, with coefficients of variation generally being < 10%. Table 1: Crystalline phase composition of fly ashes measured by quantitative x-ray diffraction and Rietveld refinement (Mass %)

Class C Class F

BC WP MR CC LS NV BP

Quartz 10.06 11.10 9.81 6.83 16.64 16.48 14.15 Mullite 0.86 0.90 1.14 - 5.08 10.17 20.01 Anhydrite 2.80 1.84 1.01 1.61 0.97 - - Lime 1.16 1.04 0.33 - - - - Periclase 3.81 2.17 2.50 1.70 0.30 0.19 - Magnetite 1.66 2.36 1.64 2.08 1.76 2.03 1.96 Merwinite 6.98 4.19 3.98 3.66 - - - Calcio-Olivine 0.43 1.34 1.57 - - - - Ilmenite - - - 0.58 - - - β-C2S 4.50 6.30 5.75 - - - - C2AS 4.45 3.27 3.76 - - - - C3A (cubic) 5.90 5.06 6.14 - - - - C3A (orthorhombic) 2.13 2.73 2.90

- - - -

Amorphous 55.25 57.69 59.47 83.53 75.25 71.13 63.87 Total 100 100 100 100 100 100 100

Page 5: Paper-WOCAConference-TJO-03132015-Final

Table 2: Oxide composition of fly ashes measured by X-ray fluorescence (Mass %)

Class C Class F

BC WP MR CC LS NV BP

SiO2 35.44 36.08 40.08 50.75 53.97 60.48 57.98 Al2O3 17.40 18.03 20.44 15.77 20.45 22.85 27.71 Fe2O3 7.15 6.02 6.29 6.28 5.62 4.47 6.35 SO3 2.34 2.91 1.62 0.79 0.52 0.46 0.25 CaO 26.45 25.90 21.35 15.05 12.71 5.20 1.64 Na2O 1.90 1.86 1.46 3.29 0.57 2.14 0.50 MgO 5.73 5.24 4.56 4.57 2.84 1.19 1.07 K2O 0.53 0.46 0.71 2.14 1.11 1.31 2.69 P2O5 0.95 1.03 1.23 0.24 0.30 0.14 0.17 TiO2 1.19 1.34 1.42 0.61 1.29 1.16 1.38 Total 99.08 98.87 99.16 99.49 99.38 99.40 99.74

3.0 QUANTIFYING THE NETWORK RATIO AS A FLY ASH DESCRIPTOR Before discussion of fly ash characterization, it is expedient to first explain the rationale behind the approach taken in this study. The majority of fly ashes are made up primarily of amorphous material, and so for the time being we confine our discussion to characterization of glass phases. It has long been recognized that calcium oxide content is linked to fly ash glass structure, as demonstrated by the marked differences between Class C and Class F fly ashes.22 While early studies correlating this with structure use simply the bulk CaO content,23 more recently the net content of network modifying elements has been considered.20,24 Silicate glasses are typically considered as a random network structure made up of silicate tetrahedra linked at each corner to exactly one other silicate tetrahedron.25,26 This structure can than incorporate other metal oxides, which contribute cations to the network. Low valence metal cations, such as calcium, potassium, magnesium, and sodium, are termed network formers, and exist in the interstitial volume. These destabilize the glass network by breaking up the structure through the creation of non-bridging oxygens for charge balance, as shown in Figure 2a.27 However, in alkali-alkaline earth silicate glasses, the alkaline earth elements are bound more tightly than alkalis, thus improving the chemical durability of the glass.26 The glass present in fly ash is typically an alkali-alkaline earth aluminosilicate glass. Aluminum is known to function as a network former for sufficiently high alkali/alkaline earth network modifier content,26,27 as is the case for the majority of fly ashes. Aluminum incorporation at tetrahedral sites normally occupied by silicon results in a net charge of -1 on the tetrahedron, which must be balanced by a cation in the interstices. This allows for the incorporation of network modifiers without the creation of additional non-bridging oxygens, as shown in Figure 2b.

Page 6: Paper-WOCAConference-TJO-03132015-Final

(a)

(b) Figure 2: An illustration of (a) how network modifiers create non-bridging oxygens, thus disconnecting the silicate tetrahedra in the glass structure, and (b) how aluminum incorporation in the network reduces non-bridging oxygens. For a single parameter to be descriptive for the range of glass compositions present in fly ash, it must take into account both network modifiers and network formers. In the past this has been accomplished by quantifying the number of non-bridging oxygens. It should be realized that this is somewhat of an oversimplification, as the field strength and coordination state of the constituent ions also may play a role, as may the formation of calcium-aluminate glass. In particular, the aluminum ion is known to alter its coordination state depending on other structural factors, and may exist in an octahedrally coordinated state, or even act as a network modifier for high [Al2O3/(M2O+2*M’O)] ratio, where M and M’ correspond to alkali and alkaline earth cations respectively.25 However, non-bridging oxygen structural models have found successful application for binary alkali glasses,26 and show good agreement with data related to structure as reported later in this study. The network ratio, as defined herein, can be calculated using equation 1 below:

𝑁𝑟 = 2 × (𝑛𝐶𝑎 + 𝑛𝑀𝑔) + 𝑛𝐾 + 𝑛𝑁𝑎 − 𝑛𝐴𝑙

𝑛𝑆𝑖 + 𝑛𝐴𝑙 𝑓𝑜𝑟 [

𝐴𝑙2𝑂3

𝑀2𝑂 + 2 × 𝑀’𝑂] < 1

(1)

𝑁𝑟 = 3 × |2 × (𝑛𝐶𝑎 + 𝑛𝑀𝑔) + 𝑛𝐾 + 𝑛𝑁𝑎 − 𝑛𝐴𝑙|

𝑛𝑆𝑖 + 𝑛𝐴𝑙 𝑓𝑜𝑟 [

𝐴𝑙2𝑂3

𝑀2𝑂 + 2 × 𝑀’𝑂] > 1

Where 𝑁𝑟 is network ratio, 𝑛𝐶𝑎, 𝑛𝑀𝑔, 𝑛𝐾, 𝑛𝑁𝑎, 𝑛𝐴𝑙, and 𝑛𝑆𝑖 are the molar concentrations

of calcium, magnesium, potassium, sodium, aluminum, and silicon respectively, and

𝑀2𝑂 + 2 × 𝑀’𝑂 represents the net charge of cations in the interstices of the glass which must be balanced by formation of either aluminum tetrahedra or non-bridging oxygen (M and M’ correspond to alkali and alkaline earth cations respectively).

Page 7: Paper-WOCAConference-TJO-03132015-Final

The above formulation assumes complete compensation of charge by aluminum tetrahedra prior to the formation of non-bridging oxygens for aluminum-modifier ratios less than one, and the formation of three moles of non-bridging oxygens for each mole of aluminum incorporated in excess of the aluminum-modifier ratio of one. As written, the network ratio represents the number of non-bridging oxygen per tetrahedral unit in

the glass structure (i.e. 𝑁𝑟 = 4 implies tetrahedral monomers, whereas 𝑁𝑟 = 0 implies complete connectivity). It should be noted that this represents the connectivity averaged across the entire glass, where in reality the connectivity of each individual tetrahedral unit may range from 0-4. A simpler formulation of the network ratio may also be employed:

𝑁𝑟 = 𝑛𝐶𝑎 + 𝑛𝑀𝑔 + 𝑛𝐾 + 𝑛𝑁𝑎

𝑛𝑆𝑖 + 𝑛𝐴𝑙 (2)

This ratio is developed from the same principles as equation 1, but omits correction for network modifying aluminum. This formulation is less sensitive to small changes in glass composition, and thus will be preferable for describing local structure as discussed later. A disadvantage of this method of calculation is the lack of a direct structural analog such as non-bridging oxygens per tetrahedral unit. In all following discussion relating to glass characterization, the network ratio, calculated from equation 1 using amorphous molar concentrations obtained by taking the difference of XRF (total) and XRD (crystalline) elemental compositions, will be used to correlate with data relating to structure and reactivity. 4.0 EXPERIMENTAL RESULTS AND DISCUSSION 4.1 Characterization of Fly Ashes As a point of reference, Figure 3a provides a graphical representation of the calcium oxide content of each ash from XRF, with the boundary between Class C and Class F demarcated at 20%. By comparison, network ratio follows a similar trend from high to low calcium ash (Figure 3b). Surface area however, shows no dependence on ash composition (Figure 3c), though it does exhibit marked similarities with loss on ignition as determined from thermal-gravimetric analysis (Figure 4a). This is in line with data from the literature, which report approximately 0.7-0.8 m2/g of surface area contributed by the inorganic fraction of a given fly ash, the remainder being due to unburned carbon.28 Loss on ignition is typically attributed solely to unburned carbon, and this is supported by the presence of its characteristically broad high temperature decomposition peak in DTG (occurring from 500-850°C),29 and the absence of calcite in the XRD data, which also would decompose in this region. The attribution of excess surface area to unburned carbon has two major implications. The first relates to compatibility of the fly ashes with chemical admixtures. High surface

Page 8: Paper-WOCAConference-TJO-03132015-Final

area unburned carbon has the capacity to strongly adsorb molecules, similar to activated carbon used in water treatment, and is known to particularly affect air entraining admixtures.28 The second relates to fly ash’s capacity to function as a mineral filler, which has been noted to exhibit strong dependence on surface area even for inert filler materials.30 The true surface area of inorganic filler particles in these fly ashes is on the order of 0.65 m2/g, determined from an average of ashes WP, CC, and LS which exhibit no decomposition in the 500-850°C range. This is nearly equivalent to the area of the cement, which is 0.636 m2/g. Given this, little if any acceleratory effect with regard to reactivity is expected to be provided as a result of surface area of each ash, with any differences in kinetic behavior arising from reactivity alone (w/c having negligible effect)30,31.

(a)

(b) (c)

Figure 3: Ash properties (a) calcium oxide content, (b) network ratio, and (c) BET specific surface area.

(a)

(b) (c)

Figure 4: (a) Loss on ignition, (b) loss on ignition, as related to specific surface area, and (c) glass transition temperature, determined from DTA heat flow by the slope-intercept method.26 To validate network ratio as a descriptive parameter for the quantification of reactivity via atomic structure, it is compared in Figure 5 to three different measurable properties which are direct indicators of atomic structure. Comparison of ash composition to the broad amorphous XRD peak and ash density was first suggested by Diamond,22,23 who

Page 9: Paper-WOCAConference-TJO-03132015-Final

showed a correlation between bulk CaO content and peak position. Peak position is a good indicator of glass structure, as it represents the short range order of the glass,20 with higher angles representing lower average interatomic spacing, and thus a more compact atomic structure. This is shown here by an increase in peak angle with increasing network ratio, indicating more compact structures for the more reactive high calcium ashes. Compact structure at high network ratio is supported by ash density, which shows a similar correlation. This may be due to both (a) shrinkage of interstices during glass formation due to smaller, high field-strength cations, and (b) associated gains dependent on atomic weight from incorporation of additional ions. Strength of both density and XRD peak correlations confirms that the densification occurs at least in part by structural means.

(a)

(b) (c)

Figure 5: Network ratio correlated with (a) amorphous XRD peak, (b) ash density, and (c) glass transition temperature. Glass transition temperature (Tg) serves as a final checkpoint for network ratio, providing a slightly less robust correlation. Glass transition is analogous to melting, and corresponds to a viscosity of 1012 Pa·s, and a change in heat capacity of the material.26 Weaker correlation is expected, as this parameter depends more on chemical durability and the conditions under which the glass was formed than either amorphous XRD peak or density, and will not be dictated entirely by structural characteristics of the glass. A direct investigation of the amorphous phases was undertaken using SEM-EDS backscatter and element map images (Figure 6b, c). Qualitatively, these provide a good illustration of the compositional heterogeneity of fly ash. Such images were acquired for all seven ashes using two fields of view to account for variability. Element maps were then used for manual clustering, and point compositions were acquired in each cluster for the elements expected from the bulk XRF composition. Multispectral image segmentation using both the backscatter images and element maps after Chancey et al.9 was carried out using automated clustering, to provide compositionally distinct groups as shown in Figure 6c. These clusters were assigned compositions based on coinciding element point map data. Cluster analysis was only carried out for three Class F ashes and one Class C ash, as the remaining ashes were too heterogenous for their cluster compositions to be resolved with respect to point mapped element data.

Page 10: Paper-WOCAConference-TJO-03132015-Final

(a) (b)

(c)

Figure 6: Scanning electron microscopy data for fly ash NV: (a) backscatter image, (b) X-ray dispersive spectroscopy element map image overlay, where calcium is shown as red, silicon as green, and aluminum as blue, and (c) multispectral image segmentation mask, key: High Alkali Glass Medium Alkali Glass Low Alkali Glass Quartz High Calcium Glass Medium Calcium Glass Low Calcium Glass

Page 11: Paper-WOCAConference-TJO-03132015-Final

(a)

(b)

Figure 7: Ternary diagrams showing (a) manually determined clusters from element maps and (b) clusters as determined by multispectral image analysis software. Red symbols indicate calcium rich phases, green indicate silicon rich phases, and blue indicate aluminum rich phases. A comparison of manual clusters to those generated by multispectral grouping is shown in Figure 7. Fairly good agreement was achieved for the three ashes analyzed in this way. Amorphous phase makeup is shown in Figure 8 b and c. For low calcium ashes (BP, NV, LS), the amount of high and medium calcium glass appears relatively fixed, with low calcium glass content rising with network modifier. Likewise, the alkali aluminosilicate contents tend to decrease with increasing network ratio as they incorporate calcium. By contrast, the high calcium Class C fly ash (MR) has a roughly equal distribution among the alkali-alkaline earth aluminosilicates, and among the alkali aluminosilicates, pointing to a decreased compositional heterogeneity, despite the drastic increase in observed spatial heterogeneity in the Class C segmenetations (phase separation occurs on a much smaller scale). These are likely effects produced by (a) the source coal and (b) the burning conditions, namely the cooling rate. As such factors are unknown for these ashes, no more can be said on the subject. With known amorphous phase compositions and fractions in hand, it is possible to calculate network ratio using a “bottom up” approach. This is done by taking an average of the network ratios of individual compositional ranges weighted by the mass fraction of that range. Network ratio calculated thusly is compared to that calculated from bulk XRF-XRD measurements in Figure 8c. Good agreement is seen for the lowest calcium ashes, though less so with increasing network ratio. This is due to the inability of the SEM technique to accurately resolve all crystalline phases, which may be microcrystalline, cryptocrystalline, or partially/wholly embedded in the amorphous particles,32 and is an inherent limitation to this method of analysis. However, close match between the multiple methods of calculation for ashes in which this problem is minimized (i.e. low calcium) provides further validation of the network ratio parameter.

SiO2

CaO Al2O

3

SiO2

CaO Al2O

3

Page 12: Paper-WOCAConference-TJO-03132015-Final

(a)

(b) (c)

Figure 8: Glass compositions for (a) calcium aluminosilicate glasses and (b) alkali aluminosilicate glasses, as a percentage of total amorphous material, and (c) a comparison between network ratio as a weighted average of the segmented glass categories and network ratio as calculated from XRF-XRD data.2

4.2 Evaluation of Reactivity and Performance Figure 9 a-g shows heat flow profiles for fly ash-cement blends. Kinetic curves look remarkably similar for all ashes, especially at low replacement levels. The primary difference observed is at high replacement levels, and can be attributed to disruption of aluminate-sulfate balance by tricalcium aluminate and/or anhydrite contained in the fly ash.7 Such similar kinetic behavior is uncharacteristic of a high surface area filler material. This can be explained by the earlier observation that the majority of the fly ash surface area may arise from unburned carbon. This would explain the nearly identical kinetic behavior, assuming unburned carbon is in no way actively involved in early reactions. An alternate explanation would be that fly ash particles themselves do not function as preferred nucleation sites the way other materials have been known to.30 However, this is not supported by microscopy data which observe fly ash particles do serve as sites for nucleation and growth of CSH at early ages.8,33 In any case, surface area effects are clearly not at play, as evidenced by the lack of dependence of the heat at peak parameter (a typical indicator of kinetic acceleration) on area multiplier as defined in equation 3 (Figure 9h).

𝐴𝑀 = 1 + 𝑟 ∙ 𝑆𝑆𝐴𝑓𝑖𝑙𝑙𝑒𝑟

(100 − 𝑟) ∙ 𝑆𝑆𝐴𝑐𝑒𝑚𝑒𝑛𝑡 (3)30

Where 𝐴𝑀 is area multiplier, 𝑟 is replacement level, and 𝑆𝑆𝐴𝑓𝑖𝑙𝑙𝑒𝑟 and 𝑆𝑆𝐴𝑐𝑒𝑚𝑒𝑛𝑡 are the

specific surface areas of the mineral filler and the cement respectively.

2 Here, the alternate formulation for network ratio found in equation 2 is used, as it was found that compensating for charge with

aluminum as a network modifier resulted in overestimations of non-bridging oxygens for local point mapped compositions. This problem arises due to the high abundance of low calcium glass, and the difficulty of resolving the cryptocrystalline phase mullite from amorphous material using SEM imaging combined with point mapped x-ray data. It should be noted that this formulation of network ratio correlates nearly as well with the structural parameters shown in Figure 5.

Page 13: Paper-WOCAConference-TJO-03132015-Final

(a)

(b) (c) (d)

(e)

(f) (g) (h)

Figure 9: Calorimetry data for (a) ash BC, (b) ash WP, (c) ash MR, (d) ash CC, (e) ash LS, (f) ash NV, and (g) ash BP (in order of decreasing CaO content), and (h) heat at peak parameter plotted against area multiplier. Initial and final set times for varying replacement levels of the seven fly ashes are shown in Figure 10. It is well known that fly ashes retard set times, as observed for these ashes.6,34 Both initial and final set time show a linear trend with replacement level, thus allowing for the rough prediction of set time within ± 1 hour. Various studies have investigated links between set time and chemical shrinkage or heat evolution, and correlations have been found to the minimum heat of the induction period,35 as well as the time of this minimum heat.6 Using these data, weak correlations of set time to these two parameters are found for 10% fly ash replacement mixtures (R2 values of 0.621 and 0.759 respectively), but these correlations break down at higher replacement levels and are not able to capture the influence of fly ash characteristics on set. In light of the fact that set time is dependent upon initial porosity (w/c) and solid-to-solid distance, it should be noted that it is inappropriate to attempt such correlations with chemical shrinkage or heat flow, which are dependent on extent of reaction.36 As such, no conclusions can be drawn as to the link between fly ash properties and set at this time. To obtain a clear picture of the effect of glass reactivity on extent of reaction, network ratio is compared to cumulative heat release at 7 days for the 50% ash systems, which will exhibit the strongest influence due to fly ash character. This is shown in Figure 11b, and demonstrates excellent correlation between network ratio and heat release, providing a direct link between structure and reactivity for the fly ashes. This correlation does not hold at low replacement levels, and further work is needed to deconvolute the transitional kinetic response for systems which fall in between cement dominated and fly ash dominated reaction regimes.

Page 14: Paper-WOCAConference-TJO-03132015-Final

(a)

(b)

Figure 10: (a) Initial and (b) Final set time as a function of fly ash replacement level. Strength-heat correlation is shown for the fly ash-cement blends in Figure 11a. Such correlations have been reported in the literature, and demonstrate that strength is developed as a direct result of heat of hydration (water normalized), with the best correlations obtained for plain cement or inert filler systems.37–39 Reactive fillers such as fly ash are expected to assert linear trends, but with slopes whose values depend on the reactivity of a given fly ash. This is evidenced by the inverse correlation of network ratio with strength-heat slope (i.e. rate of strength gain with reaction) for a given ash (Figure 11c), implying that low calcium (i.e., silicon rich) glasses contribute more effectively to strength development at a given level of reaction, likely due to the formation of a C-S-H hydrate with lower Ca/Si (molar ratio).40

(a)

(b) (c)

Figure 11: (a) Heat-strength correlation, with a linear fit and 10% error bounds for the plain paste system, and network ratio linked to (b) reactivity by correlation with cumulative heat at 7 days (i.e. extent of reaction), and (c) property development by correlation with heat-strength slope.

Page 15: Paper-WOCAConference-TJO-03132015-Final

5.0 SUMMARY AND CONCLUSIONS This study describes a general means by which the composition of a fly ash can be linked to its structure, reactivity, and performance as a SCM used to replace cement in the concrete binder. The fly ash composition is applied to assess the network ratio, which has its roots in structural features of aluminosilicate glasses. It is shown that the network ratio can be calculated by multiple means, and is correlated to well recognized indicators of glass structure such as density and glass transition temperature. Furthermore it is shown that the assessed fly ashes have minimal impact on hydration kinetics due to the similarity of their specific surface area to cement, i.e., they appear to behave loosely as inert fillers at early ages. For mixtures containing 50% fly ash (by mass), the extent of reaction assessed in terms of heat release is linked to the network ratio for fly ash containing systems, demonstrating that the parameter is not only valid for prediction of glass structure, but reactivity as well. For the same mixtures, network ratio is linked to rate of compressive strength development. This may suggest that beyond a specific OPC replacement level, the mixture’s behavior is dominantly influenced by the fly ash component, and less so by the OPC phase. The network ratio is a comprehensive parameter to link composition, structure, reactivity, and property development – across both Class C and Class F designations. Overall, the methods developed in this study provide a basis for systematic characterization and use of fly ash to promote its widespread and informed use. ACKNOWLEDGEMENTS The authors acknowledge full financial support for this research provisioned by the University of California, Los Angeles (UCLA) and the COMAX consortium. The authors would also like to acknowledge the generous provision of materials by U.S. Concrete, Lehigh Cement, and Headwaters Incorporated. The contents of this paper reflect the views and opinions of the authors, who are responsible for the accuracy of the datasets presented herein. This research was conducted in the Laboratory for the Chemistry of Construction Materials (LC2) and the Molecular Instrumentation Center (MIC) at the University of California, Los Angeles (UCLA), and the Engineering Laboratory at the National Institute of Standards and Technology (NIST). As such, the authors gratefully acknowledge the support that has made these laboratories and their operations possible. The authors would like to gratefully acknowledge Dr. Jeffrey Bullard and Dr. Paul Stutzman, for their guidance and support. The last author would also like to acknowledge discretionary support for this research provided by the Rice Endowed Chair in Materials Science.

Page 16: Paper-WOCAConference-TJO-03132015-Final

REFERENCES [1] Worrell, E., Price, L., Martin, N., Hendriks, C., Meida, L. O. Carbon Dioxide

Emissions from the Global Cement Industry 1. Annu. Rev. Energy Environ. 2001, 26, 303–329.

[2] Mehta, P. K. Sustainable Cements and Concrete for the Climate Change Era–a Review. In Second international conference on sustainable construction materials and technologies. Italy: Coventry University and The University of Wisconsin Milwaukee Centre for By-products Utilization, 2010.

[3] Mehta, P. K. High-Performance, High-Volume Fly Ash Concrete for Sustainable Development. In Proceedings of the international workshop on sustainable development and concrete technology, 2004, pp 3–14.

[4] Juenger, M. C. G., Winnefeld, F., Provis, J. L., Ideker, J. H. Advances in Alternative Cementitious Binders. Cem. Concr. Res. 2011, 41, 1232–1243.

[5] Bentz, D. P. Powder Additions to Mitigate Retardation in High-Volume Fly Ash Mixtures. ACI Mater. J. 2010, 107.

[6] Bentz, D. P., Ferraris, C. F. Rheology and Setting of High Volume Fly Ash Mixtures. Cem. Concr. Compos. 2010, 32, 265–270.

[7] Bentz, D. P., Ferraris, C. F., De la Varga, I., Peltz, M. A., Winpigler, J. Mixture Proportioning Options for Improving High Volume Fly Ash Concretes. Int. J. Pavement Res. Technol. 2010, 3, 234–240.

[8] Chancey, R. T. Characterization of Crystalline and Amorphous Phases and Respective Reactivities in a Class F Fly Ash, University of Texas at Austin, 2008.

[9] Chancey, R. T., Stutzman, P., Juenger, M. C., Fowler, D. W. Comprehensive Phase Characterization of Crystalline and Amorphous Phases of a Class F Fly Ash. Cem. Concr. Res. 2010, 40, 146–156.

[10] ASTM Standards Website. Available at: http://www.astm.org/Standards/. Accessed on 3/9/2015 at 3:03 P.m. PST.

[11] Mehta, P. K. Influence of Fly Ash Characteristics on the Strength of Portland-Fly Ash Mixtures. Cem. Concr. Res. 1985, 15, 669–674.

[12] Berry, E. E., Hemmings, R. T., Cornelius, B. J. Mechanisms of Hydration Reactions in High Volume Fly Ash Pastes and Mortars. Cem. Concr. Compos. 1990, 12, 253–261.

[13] Papadakis, V. G. Effect of Fly Ash on Portland Cement Systems: Part I. Low-Calcium Fly Ash. Cem. Concr. Res. 1999, 29, 1727–1736.

[14] Papadakis, V. G. Effect of Fly Ash on Portland Cement Systems: Part II. High-Calcium Fly Ash. Cem. Concr. Res. 2000, 30, 1647–1654.

[15] Lam, L.; Wong, Y. L.; Poon, C. S. Degree of Hydration and Gel/space Ratio of High-Volume Fly Ash/cement Systems. Cem. Concr. Res. 2000, 30, 747–756.

[16] Sakai, E., Miyahara, S., Ohsawa, S., Lee, S.-H., Daimon, M. Hydration of Fly Ash Cement. Cem. Concr. Res. 2005, 35, 1135–1140.

[17] Roy, D. M., Luke, K., Diamond, S. Characterization of Fly Ash and Its Reactions in Concrete. In MRS Proceedings, Cambridge Univ Press, 1984, Vol. 43, p 3.

[18] Pietersen, H. S. Reactivity of Fly Ash and Slag in Cement, TU Delft, Delft University of Technology, 1993.

Page 17: Paper-WOCAConference-TJO-03132015-Final

[19] Aughenbaugh, K. L., Chancey, R. T., Stutzman, P., Juenger, M. C., Fowler, D. W. An Examination of the Reactivity of Fly Ash in Cementitious Pore Solutions. Mater. Struct. 2013, 46, 869–880.

[20] Snellings, R. Surface Chemistry of Calcium Aluminosilicate Glasses. J. Am. Ceram. Soc. 2015, 98, 303–314.

[21] Janoušek, V., Farrow, C. M., Erban, V. Interpretation of Whole-Rock Geochemical Data in Igneous Geochemistry: Introducing Geochemical Data Toolkit [GCDkit]. J. Petrol. 2006, 47, 1255–1259.

[22] Diamond, S. The Utilization of Flyash. Cem. Concr. Res. 1984, 14, 455–462. [23] Diamond, S. On the Glass Present in Low-Calcium and in High-Calcium Flyashes.

Cem. Concr. Res. 1983, 13, 459–464. [24] Oh, J. E., Jun, Y., Jeong, Y., Monteiro, P. J. The Importance of the Network-

Modifying Element Content in Fly Ash as a Simple Measure to Predict Its Strength Potential for Alkali-Activation. Cem. Concr. Compos. 2015, 57, 44–54.

[25] Varshneya, A. K. Fundamentals of Inorganic Glasses, Elsevier, 1993. [26] Shelby, J. E. Introduction to Glass Science and Technology, Royal Society of

Chemistry, 2005. [27] Hemmings, R. T., Berry, E. E. On the Glass in Coal Fly Ashes: Recent Advances.

In MRS Proceedings, Cambridge Univ Press, 1987, Vol. 113, p 3. [28] Külaots, I., Hurt, R. H., Suuberg, E. M. Size Distribution of Unburned Carbon in

Coal Fly Ash and Its Implications. Fuel 2004, 83, 223–230. [29] Biernacki, J. J., Mogulla, N. R., Bollig, J., Bijou, M. Kinetics of Carbon Oxidation

from Fly Ash. Fuel 2010, 89, 1077–1086. [30] Oey, T., Kumar, A., Bullard, J. W., Neithalath, N., Sant, G. The Filler Effect: The

Influence of Filler Content and Surface Area on Cementitious Reaction Rates. J. Am. Ceram. Soc. 2013, 96, 1978–1990.

[31] Thomas, J. J. A New Approach to Modeling the Nucleation and Growth Kinetics of Tricalcium Silicate Hydration. J. Am. Ceram. Soc. 2007, 90, 3282–3288.

[32] Diamond, S. Particle Morphologies in Fly Ash. Cem. Concr. Res. 1986, 16, 569–579.

[33] Berry, E. E., Hemmings, R. T., Zhang, M.-H., Cornelius, B. J., Golden, D. M. Hydration in High-Volume Fly Ash Concrete Binders. ACI Mater. J. 1994, 91.

[34] Bentz, D. P., Sato, T., De la Varga, I., Weiss, W. J. Fine Limestone Additions to Regulate Setting in High Volume Fly Ash Mixtures. Cem. Concr. Compos. 2012, 34, 11–17.

[35] Oey, T., Stoian, J., Li, J., Vong, C., Balonis, M., Kumar, A., Franke, W., Sant, G. Comparison of Ca [NO 3] 2 and CaCl 2 Admixtures on Reaction, Setting, and Strength Evolutions in Plain and Blended Cementing Formulations. J. Mater. Civ. Eng. 2014.

[36] Sant, G., Ferraris, C. F., Weiss, J. Rheological Properties of Cement Pastes: A Discussion of Structure Formation and Mechanical Property Development. Cem. Concr. Res. 2008, 38, 1286–1296.

[37] Bentz, D. P., Durán-Herrera, A., Galvez-Moreno, D. Comparison of ASTM C311 Strength Activity Index Testing vs. Testing Based on Constant Volumetric Proportions. J ASTM Int 2012, 9, 7.

Page 18: Paper-WOCAConference-TJO-03132015-Final

[38] Bentz, D. P., Barrett, T., De la Varga, I., Weiss, W. J. Relating Compressive Strength to Heat Release in Mortars. Adv Civ Eng Mater 2012, 1, 14.

[39] Kumar, A., Oey, T., Kim, S., Thomas, D., Badran, S., Li, J., Fernandes, F., Neithalath, N., Sant, G. Simple Methods to Estimate the Influence of Limestone Fillers on Reaction and Property Evolution in Cementitious Materials. Cem. Concr. Compos. 2013, 42, 20–29.

[40] Qomi, M. A., Krakowiak, K. J., Bauchy, M., Stewart, K. L., Shahsavari, R., Jagannathan, D., Brommer, D. B., Baronnet, A., Buehler, M. J., Yip, S., et al. Combinatorial Molecular Optimization of Cement Hydrates. Nat. Commun. 2014, 5.