on a new hydraulic binder from stainless steel onverter slag

Upload: saeedhoseini

Post on 03-Apr-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    1/11

    Advances in Cement Research, 2013, 25(1), 2131

    http://dx.doi.org/10.1680/adcr.12.00031

    Paper 1200031

    Received 09/05/2012; revised 02/08/2012; accepted 10/08/2012

    ICE Publishing: All rights reserved

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

    On a new hydraulic binder fromstainless steel converter slagYiannis PontikesResearcher, Centre for High Temperature Processes and SustainableMaterials Management, Department of Metallurgy and MaterialsEngineering, KU Leuven, Belgium

    Lubica KriskovaResearcher, Centre for High Temperature Processes and SustainableMaterials Management, Department of Metallurgy and MaterialsEngineering, KU Leuven, Belgium

    Ozlem Cizer

    Researcher, Building Materials and Building Technology Division,Department of Civil Engineering, KU Leuven, Belgium

    Peter Tom JonesIOF Research Manager, Centre for High Temperature Processes andSustainable Materials Management, Department of Metallurgy andMaterials Engineering, KU Leuven, Belgium

    Bart BlanpainProfessor, Centre for High Temperature Processes and SustainableMaterials Management, Department of Metallurgy and MaterialsEngineering, KU Leuven, Belgium

    The aim of this work was to investigate the hydraulic behaviour of a stainless steel converter slag after changing

    its chemical composition and cooling path. The target slag was designed to resemble ground granulated blast-

    furnace slag (GGBFS). A synthetic slag with a chemical composition close to stainless steel converter slags was

    mixed with 22, 30 and 38 wt% fly ash (FA) from lignite combustion, heated up to 15508C and then granulated by

    quenching in water; the solidified new slags were named FA22, FA30 and FA38 respectively. Quantitative X-ray

    diffraction on FA22 revealed that the amorphous phase was approximately 40 wt%, the rest being bredigite and

    merwinite. For FA addition of 30 wt% or more, the amorphous phase reached almost 100 wt%. The resulting slags

    showed significant hydraulic activity when mixed with sodium-based activators, with C-S-H, hydrotalcite and

    hydrogarnet being the main hydration products formed. The calorimetric behaviour and the mechanical properties

    of blended cements with 30 wt% FA30 and FA38 were comparable to a blended cement with GGBFS. Assuming

    that FA addition will take place during the liquid state of the slag, the proposed process can result in a new

    hydraulic binder.

    IntroductionGround granulated blast-furnace slag (GGBFS) is one of the

    most widely used supplementary cementitious materials in

    blended Portland cements. It is produced by water granulation of

    a blast-furnace slag, forming a material in which the principal

    component is a calcium magnesium aluminosilicate glass (Wangand Scrivener, 2003). GGBFS has latent hydraulic properties,

    implying that the slag reacts with water to give a cementitious

    material once activated in the presence of Portland cement, lime

    or alkalis such as caustic soda, sodium carbonate or sulfates of

    alkali, calcium or magnesium (Lang, 2002). The hydration of

    GGBFS is slow when compared with Portland cement clinker,

    resulting in lower strength gain at early stages and higher

    strength gain at later stages (Taylor, 1990). The hydration of slag

    proceeds by way of dissolution of slag particles followed by

    precipitation of hydrated phases from the supersaturated pore

    solution. Since the dissolution can be accelerated at high pH

    values in the pore solution, there is a tendency to use Portland

    cement clinker with a higher content of water-soluble alkalis to

    produce blended cement containing GGBFS (Bellmann and

    Stark, 2009). In terms of applications, a number of studies have

    confirmed that GGBFS is an economical, environmentally

    friendly and highly chemically resistant component of building

    materials (Mozgawa and Deja, 2009). The European standard

    EN 197-1: 2000 (CEN, 2000) reflects the above: CEM III/C can

    contain up to 95 wt% GGBFS, which delivers a hydraulic binder

    with a particularly low carbon dioxide footprint.

    Stainless steel slags are typically used as aggregates and only afew, higher value applications, such as fertiliser, are practised

    worldwide (Engstrom et al., 2011). Considering that GGBFS is a

    material with a relatively high added value compared with other

    slags, it is worthwhile to investigate if stainless steel slags could

    be converted to GGBFS-like materials. In addition, the fact that

    GGBFS has been well studied provides end users with relative

    confidence regarding its behaviour and minimises the so-called

    non-technical barriers often encountered when new building

    materials are introduced (van Deventer et al., 2010).

    The aim of this work was thus to evaluate the potential of

    synthesising a material with similar properties as GGBFS, after

    modifying a stainless steel converter slag with additions of a

    silicon, aluminium-rich industrial waste (i.e. fly ash). In the

    envisaged process, the final end-product, if proven similar to

    GGBFS in terms of performance, could be applied in blended

    cements, mortars, pre-cast concrete or even inorganic polymers.

    21

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    2/11

    Experimental methodA mixture of analytical grade oxides and carbonates with an

    elemental composition close to typical stainless steel converter

    slags was mixed with 22, 30 and 38 wt% of industrially produced

    fly ash (FA) from lignite combustion (Table 1). The FA used is

    classified as type F according to ASTM C618-01 (ASTM, 2008)

    with quartz (SiO2), anorthite (CaAl2Si2O8), magnetite (Fe3O4),

    anhydrite (CaSO4) and gehlenite (Ca2Al2SiO7) identified as the

    main crystalline phases. To ensure homogeneity of the powders,

    the final compositions were mixed in a Turbula T2C mixer for at

    least 12 h.

    The resulting material was placed in a platinum crucible and

    melted in a bottom loading furnace (AGNI ELT 160-02), at

    15508C, with a heating rate of 58C/min. After an isothermal step

    at maximum temperature for 1.5 h, the melt was quenched in

    water. The final product, vitreous in nature, was dried in air and

    milled for 2 h in a bead mill (Dispermat SL-12-C1, VMA) at

    5000 rpm. The particle size distribution was determined by laser

    scattering technique (MasterSizer Micro Plus, Malvern). Each

    powder was measured three times and the average values are

    reported. The final slag mixtures with 22, 30 and 38 wt% FA are

    named FA22, FA30 and FA38 respectively and have compositions

    that fit into the range of GGBFS (Bhatty et al., 2004) (see Table

    1). An industrially produced GGBFS was also integrated in the

    research programme and used as a reference material in the studyof the hydraulic behaviour of the slag mixtures.

    The mineralogical composition and the amorphous content were

    determined by X-ray powder diffraction (XRPD, D500 Siemens)

    and Rietveld analysis using Topas Academic software. Materials

    were mixed with 10 wt% of zinc oxide and measured over a 2

    range of 10708 using CuK radiation of 40 kV and 40 mA, with

    a 0.028 step size and step time of 4 s.

    The reactivity of the synthetic slags was studied after alkali

    activation with analytical grade solutions of sodium hydroxide

    (NaOH (NH)), sodium carbonate (Na2CO3 (NC)) and sodiumsilicate, with a nominal molecular formula Na2O.3.4SiO2 (NS).

    In all cases, the sodium oxide to slag ratio was equal to 8 wt%,

    while keeping the solid to liquid equal to 1. The paste samples

    were subjected to isothermal conduction calorimetry (TAM Air

    device, TA Instruments) at 208C to monitor heat release during

    the reaction.

    To gain more insight into the reaction mechanism and reaction

    products, paste samples were prepared by mixing the slag with

    selected alkali activators (conditions as above). The pastes were

    stored in closed plastic capsules for 3, 7 and 28 days. Samples

    were subsequently crushed into powder and were vacuum dried at

    0.035 mbar for 2 h (Alpha 1-2 LD, Martin Christ), as suggested

    elsewhere (Knapen et al., 2009). Fourier transform infrared

    spectroscopy (FTIR) (Alpha spectrometer, Bruker) was employed

    to reveal bond structure information. For the measurement,

    approximately 4.5 mg of hand-ground sample was mixed with

    450 mg of KBr and compressed into pellets. Thermal analysis of

    the dried samples was performed using TGA/DSC (STA 409 PC

    Luxx1, Netzsch). The samples were heated at 58C/min in a

    continuous nitrogen gas flow up to 10008C. The microstructure of

    the hydrated product was studied by means of a scanning electron

    microscope (SEM XL30, Phillips). For this purpose, bulk samples

    of hydrated pastes were dried at 508C for 2 days.

    Finally, blended mortar samples composed of ordinary Portland

    cement (OPC, CEM I, 42.5 R) in 70 wt% and slag mixtures in

    30 wt% were prepared based on EN 196-1 (CEN, 2005). CEN

    standard sand (02 mm particle size) was used in a binder to sand

    ratio of 1:3 and in a binder to water ratio of 0 .5 by mass. The

    mortar mixtures were cast in 20 mm 3 20 mm 3 160 mm moulds

    and stored at 208C and relative humidity .95%. Compressive and

    flexural strength tests were performed using an Instron 4467. Four

    measurements for compressive strength and two for flexural

    strength per mortar sample and hydration time were performed

    and the average values and standard deviations are reported.

    Composition: wt%

    CaO SiO2 MgO Al2O3 Fe2O3 SO3 Other

    Typical GGBFS 30 50 27 40 1 10 5 15 ,1 0.62

    GGBFS 41.9 35.5 9.2 9.5 0.4 0.8 2.7

    FA 12.1 42.9 5.4 22.9 6.6 6.3 3.8

    Synthetic slag 56.7 28.4 6.5 1.3 1.1 6.0

    FA22 46.9 31.6 6.3 6.1 1.5 1.4 6.2

    FA30 43.3 32.7 6.2 7.8 1.9 1.9 6.2

    FA38 39.8 33.9 6.1 9.5 2.5 2.4 5.8

    Table 1. Chemical composition of typical GGBFS (Bhatty et al.,

    2004 and references therein) and the GGBFS, FA, synthetic slag

    and the three slag mixtures studied in this work

    22

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    3/11

    Results and discussion

    Material characterisation

    The quenched slag was glassy and dark brownish colour,

    primarily due to the presence of iron (Figure 1). The originally

    produced granules (Figure 1(a)) could be easily broken by hand

    into angular fragments due to the extensive formation of cracks

    (Figure 1(b)).

    Results from X-ray diffraction (XRD) (see Figure 2) and Rietveld

    analysis revealed that with an addition of 22 wt% FA to the

    synthetic slag, the quenched material contained approximately

    30 wt% merwinite, 30 wt% bredigite, the rest being an amor-

    phous phase. For 30 wt% and 38 wt% addition of FA, the material

    was almost completely amorphous. The industrial sample of

    GGBFS was also mainly amorphous, with merwinite identified as

    the only crystalline phase present.

    After milling, all powders showed similar particle size distribu-

    tion, with d10 , 0.5 m, d50 , 3 m andd90 , 9 m.

    Reactivity with alkalis

    Isothermal conduction calorimetry results of slags activated with

    different alkali solutions are presented in Figures 3 (a)(d). The

    effect of an activator depends mainly on its nature, dosage andcharacteristics of the activated material (Ben Haha et al., 2011a,

    2011b, 2012; Shi et al., 2006). Consequently, activation of

    different materials results in the formation of different hydration

    products with different properties (Shi and Day, 1996; Shi et al.,

    2006).

    Activation with NH resulted in the highest recorded heat release

    among all activators, for the time investigated, with the exception

    of GGBFS and activation by way of NC where the detected heat

    release was slightly higher. Moreover, NH was the only activator

    that gave a clear peak for all three slags. For FA22, the main peak

    occurred after approximately 10 h of hydration, whereas in thecase of FA30 and FA38 the peak occurred faster, implying the

    acceleration of hydration reactions. According to Shi et al.

    (2006), NH-activated slags typically have a calorimetry curve

    consisting of two peaks: the first peak, before the induction

    period, is attributed to wetting and dissolution, and the second

    peak, after the induction period, is ascribed to accelerated

    hydration. However, a double peak was clearly visible only in

    FA22; FA30 showed also two peaks but the time interval between

    was ,3 h, whereas FA38 and GGBFS reacted even faster and

    showed only a single peak and no peak respectively. These data

    suggest that the reaction kinetics are enhanced as the FA content

    in the synthetic slags increases but do not reach that of GGBFS.

    Regarding NS, a clear peak of heat release was observed for

    FA22 and FA30, unlike FA38, where there was no distinct peak

    and moreover the cumulative heat release was substantially

    smaller. This could be attributed to the slower kinetics of

    (b)

    1 cm

    (a)

    1 mm

    Figure 1. Water-quenched material (a) detailed view and (b) as

    produced

    20 30 40 50 60 70

    2 : degrees

    Intensity:arbitraryunits

    FA22

    FA30

    FA38

    GGBFS

    1

    2, 3

    1

    1

    2, 3 2, 31 1 1 1

    2

    1 ZnO2 Merwinite3 Bredigite

    Figure 2. XRD patterns of FA22, FA30, FA38 and GGBFS; 10 wt%

    zinc oxide is added as internal standard

    23

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    4/11

    NS-activated systems (Ben Haha et al., 2011a). NC, on the other

    hand, was effective only in the case of FA30 and FA38. Similarlyto NS, the peak in NC-activated samples occurred at a later time

    for an increasing FA content in the original slag (i.e. lower

    basicity, higher polymerisation in the glass matrix of the slag).

    Finally, even though both NS and NC belong to the activators

    giving a hydration curve starting with a double peak before the

    induction period and one peak after the induction period (Shi et

    al., 2006), this was not apparent in the current study.

    Activation with CH gave no peaks of heat release and resulted in

    comparatively low cumulative heats for all samples tested (Figure

    3). This reflects the slower hydration kinetics and is attributed

    mainly to the lower pH in the pore solution compared with alkali

    activation, which subsequently defines a slower dissolution rate

    for the slag (Bellmann and Stark, 2009).

    From the FA addition point of view, FA30 was the only material

    that showed clear peaks of heat release for all three sodium-based

    activators. FA30 also generated the largest amount of heat during

    the hydration when activated with every activator except CH.

    Reactivity and hydration products after activation with

    sodium hydroxide

    The XRD patterns of slags activated with NH having zinc oxide as

    internal standard are presented in Figure 4. A broad hump present

    in each non-hydrated slag in the 2 region of 25388 slightly

    diminished during the first 3 7 days of hydration and a new

    diffuse peak at about 2 29.58 appeared. This peak is assigned

    to C-S-H phase, JCPDS-ICDD # 45-1480 (Song and Jennings,

    1999). C-S-H is generally considered to be poorly crystalline but

    its crystallinity in sodium hydroxide-activated slag has already

    been reported by Shi et al. (2006). Other crystalline phases such as

    hydrotalcite (identified as Mg6Al2CO3(OH)16:

    4H2O, JCPDS-

    ICDD # 41-1428) and hydrogarnet (identified as katoite

    Ca3Al2(OH)12, JCPDS-ICDD # 24-217) were also identified.

    Interestingly, hydrogarnet was only detected in the hydrated slag

    samples in which FA was incorporated. The latter may be related

    NH NS NC CH

    Open symbols for cumulative heat release

    0

    2

    4

    6

    8

    10

    12

    14

    16

    0

    50

    100

    150

    0 10 20 30 40 50 60 70 80 90

    Rateofheatrelease:J/gperhour

    Time: h(a)

    Cumulativeheatrelease:J/g

    0

    2

    4

    6

    8

    10

    12

    14

    16

    0

    50

    100

    150

    200

    0 10 20 30 40 50 60 70 80 90

    Rateofheatrelease:J/gperhour

    Time: h(b)

    Cumulativeheatrelease:J/g

    0

    2

    4

    6

    8

    10

    12

    14

    16

    0

    50

    100

    150

    200

    0 10 20 30 40 50 60 70 80 90 100 110

    Rateo

    fheatrelease:J/gperhour

    Time: h(c)

    Cumulativeheatrelease:J/g

    0

    2

    4

    6

    8

    10

    12

    14

    16

    0

    50

    100

    150

    200

    0 10 20 30 40 50 60 70 80 90

    Rateo

    fheatrelease:J/gperhour

    Time: h(d)

    Cumulativeheatrelease:J/g

    Figure 3. Isothermal conduction calorimetry of slags activated

    with NH, NS, NC and CH: (a) FA22; (b) FA30; (c) FA38;

    (d) GGBFS

    24

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    5/11

    to the slightly different chemistry of the synthetic slags and in

    particular the higher iron content (Table 1). The main peak of

    calcite calcium carbonate (CaCO3) at 2 29.48 (JCPDS-ICDD #

    5-586) is very close to the C-S-H region and most probably calciteis present in a small amount; thermogravimetric analysis (TGA)

    and FTIR results presented later on corroborate the presence of

    carbonates. In general, the above findings are in good agreement

    with other published research (e.g. Puertas and Fernandez-Jime-

    nez, 2003; Shi et al., 2006; Song and Jennings, 1999).

    Thermogravimetric analysis and differential thermogravimetry

    (DTG) were used to monitor the hydration progress during the

    first 28 days (Figures 5 (a)(d)). For all samples activated with

    NH, the first peak observed in the DTG curves (Figures 5 (a)

    (d)) was at 851058C and is attributed to C-S-H decomposition

    (Hewlett, 1998). The intensity of this peak increased with

    increasing hydration time up to 28 days, which is an indication of

    increasing C-S-H formation (assuming no variation in the

    stoichiometry of C-S-H). Occasionally, a shoulder was detected at

    approximately 1358C, an indication of AFm type phases (e.g.

    C4AH13; see Taylor (1990)). Additional peaks are clearly ob-

    served in the DTG curves between approximately 250 and 3808C.

    Most probably, the peak at approximately 2608C is due to the

    decomposition of hydrogarnet (Passaglia and Rinaldi, 1984;

    Rivas-Mercury et al., 2008) and the main peak between 3008C

    and 3358C is due to the decomposition of hydrotalcite (Hickey et

    al., 2000; Wang and Scrivener, 1995).

    In terms of reaction kinetics, the weight loss after 28 days of

    hydration was substantial for all studied materials and comparable

    with the reference GGBFS for the same activator (Figure 5 (a)

    (d)). In detail, FA22 showed a small weight loss of 12.8 wt% after

    3 days of hydration. The reactions were very slow between day 3

    and day 7, but an acceleration was observed later on, resulting in

    a weight loss of 24.3 wt% after 28 days. The above trend is

    partially explained by the rather sluggish hydration kinetics, as

    also suggested by the calorimetry data (Figure 3). The highest

    reaction rate at early stage was observed in FA30, where a

    substantial weight loss of 19.9 wt% was recorded after 3 days of

    hydration. The reactions continued but at a decreasing rate for

    the rest of the period studied. The total weight loss after 28 days

    was 24.8 wt%. Sample FA38 reacted similarly to FA22 during the

    20 30 40 50 60 702 : degrees

    (a)

    Intensity:arbitraryunits

    Original

    3 days

    7 days

    28 days

    90 days

    6

    2 2 2

    6 5 63, 4

    1, 4

    6

    1

    1

    6 61, 2

    21 1 1

    1 ZnO

    3 CaCO2 Merwinite

    3

    4 C-S-H5 Hydrotalcite6 Hydrogarnet

    20 30 40 50 60 702 : degrees

    (b)

    Intensity:arbitraryunits

    Original

    3 days

    7 days

    28 days

    90 days

    1 ZnO2 CaCO33 C-S-H

    4 Hydrotalcite5 Hydrogarnet

    5 5 4 52, 3

    1, 31

    1

    5 51

    3 5

    1 15

    1

    20 30 40 50 60 702 : degrees

    (c)

    Intensity:arbitraryunits

    Original

    3 days

    7 days

    28 days

    90 days

    1 ZnO2 CaCO33 C-S-H

    4 Hydrotalcite5 Hydrogarnet

    5 5 4 52, 3

    1, 31

    1

    5 51

    3

    1 15

    1

    20 30 40 50 60 702 : degrees

    (d)

    Intensity:arbitraryunits

    Original

    3 days

    7 days

    28 days

    90 days

    1 ZnO2 CaCO33 C-S-H

    4 Hydrotalcite5 Merwinite

    42, 3

    1, 31

    1

    51

    31 1 1

    Figure 4. XRD patterns of alkali-activated hydrated pastes at 3,

    7, 28 and 90 days: (a) FA22 + NH; (b) FA30 + NH; (c) FA38 + NH;

    (d) GGBFS + NH

    25

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    6/11

    first 3 days. After the third day, the reactions slowed down. The

    weight loss between day 3 and day 7 of hydration was almost as

    much as that between day 7 and day 28. The total weight loss

    after 28 days of hydration was 24.3 wt%.

    Figure 6 presents the FTIR spectra of all the slags before and after

    90 days of hydration. The band at 500 cm1 is assigned to TO T

    or OTO (T Si, Al) bending vibration (McMillan, 2001)

    whereas the band at 700 cm1 is assigned to symmetric stretching

    vibration of SiO T bonds (e.g. Pechar and Rykl, 1983). The

    main band at 950 cm1 is an assemblage of symmetric SiO

    stretching vibrations of tetrahedral silicate forms with one, two,

    three and four non-bridged oxygen atoms per silicon atom (NBO/

    Si) (McMillan, 2001). The band at 1415 1430 cm1 is assigned to

    CO stretching vibration (e.g. Tatzber et al., 2007) whereas peaks

    at about 3450 cm1 and 1640 cm1 are assigned to OH stretch-

    ing and H O H bending respectively (e.g. Pechar and Rykl,

    1983). After 90 days, the main broad peak at 950 cm1 became

    significantly smaller and narrower, whereas new peaks appeared or

    were different in their intensities. In detail, the typical bands of

    C-S-H gel are detected as peaks appearing

    j at 950970 cm1, corresponding to the SiO asymmetric

    stretching bands in Q2 units

    DTG:mg/mgperC

    3 days

    7 days

    28 days

    200 400 600 800 1000

    Temperature: C(a)

    3 days

    7 days

    28 days

    100

    200 400 600 800 1000Temperature: C

    989694929088868482

    80787674

    Weight:% D

    TG:mg/mgperC

    3 days

    7 days

    28 days

    200 400 600 800 1000

    Temperature: C(b)

    3 days

    7 days

    28 days

    100

    200 400 600 800 1000Temperature: C

    98969492908886848280787674

    Weight:%

    DTG:mg/mgper

    C

    3 days

    7 days

    28 days

    200 400 600 800 1000

    Temperature: C

    (c)

    3 days

    7 days

    28 days

    100

    200 400 600 800 1000Temperature: C

    98969492908886848280787674

    Weight:%

    DTG:mg/mgper

    C

    3 days

    7 days

    28 days

    200 400 600 800 1000

    Temperature: C

    (d)

    3 days

    7 days

    28 days

    100

    200 400 600 800 1000

    Temperature: C

    98969492908886848280787674

    Weight:%

    Figure 5. DTG and TGA (insets) of hydrated pastes activated with

    NH at 3, 7 and 28 days: (a) FA22; (b) FA30; (c) FA38; (d) GGBFS

    26

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    7/11

    Transmitance:arbitraryunits

    90 days

    Original

    3500 1500 1000 500

    (a)

    90 days

    Original

    3500 1500 1000 500

    (b)

    Transmitan

    ce:arbitraryunits

    90 days

    Original

    3500 1500 1000 500

    Wave number: cm

    (c)

    1

    90 days

    Original

    3500 1500 1000 500

    Wave number: cm1

    (d)

    Figure 6. FTIR spectra of original slag samples and activated slag

    hydrated for 90 days: (a) FA22 + NH; (b) FA30 + NH;

    (c) FA38 + NH; (d) GGBFS

    27

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    8/11

    j at about 815820 cm

    1, ascribed to the SiO symmetricvibrations in Q1 units

    j in the region of 500400 cm1, notably new peaks at

    490 cm1, 450 cm1 and 422 cm1, which are associated

    with vibrations of the OSiO bonds (Mozgawa and Deja,

    2009; Ping et al., 1999; Puertas and Fernandez-Jimenez,

    2003).

    The peak at 670 cm1 is associated with the vibrations of Si O Al

    bridges (Mozgawa and Deja, 2009). Both characteristic peaks at

    about 3450 cm1 and 1640 cm1, assigned to OH and HOH

    vibrations respectively, have higher intensity in the hydrated

    samples. The presence of carbonate groups [CO3]2 is evidenced

    by peaks at about 1433 cm1: This is in agreement with the XRD

    and TGA data supporting the presence of hydrotalcite and CaCO 3:

    Microstructural analysis

    The microstructural development during hydration depended on

    the chosen activator. FA30 and FA38 behaved similarly for the

    same activator. Figure 7 shows the SEM results for FA30

    activated with NH during the first 90 days.

    The microstructure at 3 days of hydration (Figure 7(a)) was

    characteristic of C-S-H crystal growth. The original slag particles

    are covered with fibrillar crystals and a reticular network with

    distinctive bridges started to form. Porosity remained substan-tial, however; even at 7 days of hydration, a more compact

    structure had evolved (Figure 7(b)). The fibrillar C-S-H morph-

    ology was still apparent, yet the crystals were well developed,

    forming densified isles. Plate-like crystals were occasionally

    detected; based on their size and characteristic morphology, they

    are most probably AFm phases (Wang and Scrivener, 1995,

    2003). This is in line with the DTG results (Figure 5) and the

    discrete peaks detected at approximately 1358C. At 28 days of

    hydration (Figure 7(c)), former slag grains were reduced to

    particles of less than 1 m and an extensive cohesive network

    was formed in the inter-particle space. Clusters of elongated

    crystals could be seen occasionally; these are attributed to C-S-H.Similar C-S-H morphologies have also been detected elsewhere

    for blast-furnace slag finely milled and after 28 days of hydration

    (Kumar et al., 2005). At 90 days of hydration (Figure 7(d)), the

    plate-like AFm crystals were well intercalated into the dense

    matrix. Regarding hydrotalcite, the characteristic platelets were

    not detected in the hydrated microstructure, probably as the result

    of the small size (e.g. Ben Haha et al., 2011b). On the contrary,

    hydrogarnet could be distinguished more easily due to the

    characteristic trapezohedral crystals (Figure 7(e)).

    Behaviour in blended cements with OPC and comparison

    with GGBFS

    To further compare the hydraulic potential of the slags developed

    in this work with currently produced GGBFS, blended cements

    were developed with OPC and the slags or the reference GGBFS.

    Two aspects were evaluated the hydration behaviour by means

    of isothermal calorimetry and the mechanical properties.

    In terms of heat release (Figure 8), the blends with GGBFS,FA30 and FA38 behaved almost identically. All blended cements

    showed the characteristic double peak more as a shoulder than

    separate peaks. Similar calorimetry curves were reported by

    Meinhard and Lackner (2008) who investigated the hydration of

    GGBFS and OPC blends at room and elevated temperature. The

    blends showed the peak of the heat release rate at approximately

    25 h whereas, at 100 h, the evolved cumulative heat release is

    slightly above 120 J/g. Interestingly, the cement blend with FA22

    evolved heat faster, having its peak between 15 h and 25 h

    approximately. It is possible that the presence of crystals

    (merwinite, bredigite) in FA22 affected the hydration kinetics by

    providing additional nucleation sites.

    The results of flexural and compressive strength up to 90 days

    shown in Figures 9 and 10 respectively reveal that the blends

    with FA30 and FA38 behave similarly and are very close to the

    reference mixture with GGBFS. The blend with FA22 had a

    slower strength gain and a small overall strength at 90 days.

    Comparison of all blended cements with CEM I showed that

    strength gain is slower, as expected for cements with GGBFS, yet

    at 28 and 90 days there was no notable difference. With respect

    to flexural strength, no clear trend was observed as all blends

    presented comparable values for a chosen hydration day.

    Considerations for industrial implementationThe presented results demonstrate that production of a replica

    blast-furnace slag may be a viable option for the valorisation of

    secondary steelmaking slags. The required processing would take

    place after slag metal separation and some recent papers

    (Engstrom et al., 2011; Pontikes et al., 2011) clearly demonstrate

    that there is know-how to perform such an operation. This new

    slag could be produced from various waste materials that

    currently find limited or no use (e.g. high-carbon FA or secondary

    aluminas), by effectively controlling the chemistry. As a result,

    this slag becomes a product itself and not a by-product or residue.

    The latter is important as it appears to be possible to design

    tailor-made binders for specific applications (blended cement oralkali-activated). Industrial implementation would probably re-

    quire substantial investment and secured access to secondary

    resources. It is thus likely that local conditions will eventually

    dictate how industrially realistic such a process is and whether

    industrial symbiosis could occur.

    Conclusionsj Addition of fly ash (FA) to a synthetic slag close in

    composition to stainless steel converter slag resulted in a

    comparable material to blast-furnace slag in terms of

    chemistry and mineralogy.

    j The amount of crystalline phase decreased when the amount

    of FA increased and, for an addition of 30 wt% FA and above,

    the resulting material was almost completely amorphous.

    j Activation with sodium hydroxide typically resulted in

    comparatively fast and substantial heat release. Only the FA30

    sample could be activated with all three alkali activators.

    28

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    9/11

    j The slags activated with sodium hydroxide resulted in the

    formation of C-S-H, hydrotalcite, hydrogarnet and AFm

    phases as the main hydration products.

    j The calorimetric behaviour and mechanical properties of

    blended cements with FA30 and FA38 were very similar to a

    blended cement with GGBFS.

    j Production of such slags by way of hot-stage processing

    could upgrade the currently produced secondary

    (a) (b)

    (c) (d)

    (e)

    Figure 7. SEM images of FA38 hydrated for (a) 3 days, (b) 7 days,

    (c) 28 days and (d, e) 90 days. The arrows in (e) point to

    hydrogarnet crystals

    29

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    10/11

    steelmaking slags with regard to their applications anddeliver a new binder, engineered specifically for particular

    applications.

    AcknowledgementsThe authors gratefully acknowledge IWT O&O project 090594

    for financial support. Y. Pontikes and O. Cizer thank the FWO for

    post-doctoral fellowships.

    REFERENCES

    ASTM (2008) ASTM C618-01: Standard specification for coal fly

    ash and raw or calcined natural pozzolan for use as a mineral

    admixture in concrete. ASTM, West Conshohocken, PA,

    USA.

    Bellmann F and Stark J (2009) Activation of blast furnace slag by

    a new method. Cement and Concrete Research 39(8): 644

    650.

    Ben Haha M, Le Saout G, Winnefeld F and Lothenbach B

    (2011a) Influence of activator type on hydration kinetics,

    hydrate assemblage and microstructural development of alkali

    activated blast-furnace slags. Cement and Concrete Research

    41(3): 301310.

    Ben Haha M, Lothenbach B, Le Saout G and Winnefeld F

    (2011b) Influence of slag chemistry on the hydration of

    alkali-activated blast-furnace slag Part I: Effect of MgO.Cement and Concrete Research 41(9): 955963.

    Ben Haha M, Lothenbach B, Le Saout G and Winnefeld F (2012)

    Influence of slag chemistry on the hydration of alkali-

    activated blast-furnace slag Part II: Effect of Al2O3:

    Cement and Concrete Research 42(1): 74 83.

    Bhatty JI, Miller FM and Kosmatka SH (eds) (2004) Innovations in

    Portland Cement Manufacturing, 1st edn. Portland Cement

    Association, Skokie, IL, USA.

    CEN (Comite Europeen de Normalisation) (2000) EN 197-1:

    2000. Cement part 1. Composition, specifications and

    conformity criteria for common cements. CEN, Brussels,

    Belgium.CEN (2005) EN 196-1. Methods of testing cement part 1.

    Determination of strength. CEN, Brussels, Belgium.

    Engstrom F, Pontikes Y, Geysen D et al. (2011) Review: hot stage

    engineering to improve slag valorisation options. Proceedings

    of the 2nd International Slag Valorisation Symposium,

    Leuven, Belgium, pp. 231 251.

    Hewlett PC (ed.) (1998) Leas Chemistry of Cement and

    Concrete. Butterworth-Heinemann.

    Hickey L, Kloprogge JT and Frost RL (2000) The effects of

    various hydrothermal treatments on magnesiumaluminium

    hydrotalcites. Journal of Materials Science 35(17): 4347

    4355.

    Knapen E, Cizer O, Van Balen K and Van Gemert D (2009) Effect

    of free water removal from early-age hydrated cement pastes

    on thermal analysis. Construction and Building Materials

    23(11): 34313438.

    Kumar R, Kumar S, Badjena S and Mehrotra S (2005) Hydration

    0

    1

    2

    3

    4

    5

    6

    0

    20

    80

    120

    160

    0 20 40 60 80 100

    Rateofheatrelease:J/gperhour

    Time: h

    Cumulativeheatrelease:J/g140

    100

    60

    40

    07 OPC 03 GGBFS07 OPC 03 FA2207 OPC 03 FA3007 OPC 03 FA38

    Open symbols for cumulative

    Figure 8. Rate and cumulative heat release from isothermal

    calorimetry measurements for blended cements with 70 wt%

    OPC and 30 wt% GGBFS, FA22, FA30 or FA38

    0

    1

    2

    3

    45

    6

    7

    8

    9

    10

    11

    12

    0 10 20 30 40 50 60 70 80 90

    Flexuralstrength:MPa

    Hydration time: days

    OPC

    07 OPC 03 GGBFS

    07 OPC 03 FA22

    07 OPC 03 FA30

    07 OPC 03 FA38

    Figure 9. Flexural strength of blended mortars with 70 wt% OPC

    and 30 wt% GGBFS, FA22, FA30 or FA38

    0

    10

    20

    30

    40

    50

    0 10 20 30 40 50 60 70 80 90

    Compressivestrength:MPa

    Hydration time: days

    OPC

    07 OPC 03 GGBFS

    07 OPC 03 FA22

    07 OPC 03 FA30

    07 OPC 03 FA38

    Figure 10. Compressive strength of blended mortars with

    70 wt% OPC and 30 wt% GGBFS, FA22, FA30 or FA38

    30

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain

  • 7/28/2019 On a New Hydraulic Binder From Stainless Steel Onverter Slag

    11/11

    of mechanically activated granulated blast furnace slag.Metallurgical and Materials Transactions B 36(6): 873883.

    Lang E (2002) Blastfurnace Cements, 2nd edn. Spon Press, New

    York, NY, USA.

    McMillan PF (2001) Amorphous materials: vibrational

    spectroscopy. In Encyclopedia of Materials: Science and

    Technology. (Buschow KHJ, Chan RW, Flemings MC (eds)).

    Elsevier, pp. 251256.

    Meinhard K and Lackner R (2008) Multi-phase hydration model

    for prediction of hydration-heat release of blended cements.

    Cement and Concrete Research 38(6): 794802.

    Mozgawa W and Deja J (2009) Spectroscopic studies of alkaline

    activated slag geopolymers. Journal of Molecular Structure

    924926: 434441.

    Passaglia E and Rinaldi R (1984) Katoite, a new member of the

    Ca3Al2(SiO4)3Ca3Al2(SiO4)3(OH)12 series and a new

    nomenclature for the hydrogrossular group of minerals.

    Bulletin de Mineralogie 107(5): 605618.

    Pechar F and Rykl D (1983) Study of the complex vibrational

    spectra of natural zeolite mordenites. Zeolites 3(4): 329332.

    Ping Y, Kirkpatrick RJ, Brent P, McMillan PF and Cong X (1999)

    Structure of calcium silicate hydrate (C-S-H): Near-, mid-,

    and far-infrared spectroscopy. Journal of the American

    Ceramic Society 82(3): 742748.

    Pontikes Y, Kriskova L, Wang X et al. (2011) Additions of

    industrial residues for hot stage engineering of stainless steelslags. Proceedings of the 2nd International Slag Valorisation

    Symposium, Leuven, Belgium, pp. 313 326.

    Puertas F and Fernandez-Jimenez A (2003) Mineralogical and

    microstructural characterisation of alkali-activated fly

    ash/slag pastes. Cement and Concrete Composites 25(3):287292.

    Rivas-Mercury JM, Pena P, de Aza AH and Turrillas X (2008)

    Dehydration of Ca3Al2(SiO4)y(OH)4(3-y) (0 , y , 0.176)

    studied by neutron thermodiffractometry. Journal of the

    European Ceramic Society 28(9): 17371748.

    Shi C and Day RL (1996) Selectivity of alkaline activators for the

    activation of slags. Cement, Concrete and Aggregates 18(1):

    814.

    Shi C, Krivenko PK and Roy D (2006) Alkali-Activated

    Cements and Concretes. Taylor & Francis, New York, NY,

    USA.

    Song S and Jennings HM (1999) Pore solution chemistry of

    alkali-activated ground granulated blast-furnace slag. Cement

    and Concrete Research 29(2): 159170.

    Tatzber M, Stemmer M, Spiegel H et al. (2007) An alternative

    method to measure carbonate in soils by FT-IR spectroscopy.

    Environmental Chemistry Letters 5(1): 912.

    Taylor HFW (ed.) (1990) Cement Chemistry, 2nd edn. Academic

    Press, London, UK.

    van Deventer J, Provis J, Duxson P and Brice D (2010) Chemical

    research and climate change as drivers in the commercial

    adoption of alkali activated materials. Waste and Biomass

    Valorization 1(1): 145155.

    Wang SD and Scrivener KL (1995) Hydration products of alkali

    activated slag cement. Cement and Concrete Research 25(3):561571.

    Wang SD and Scrivener KL (2003) 29Si and 27Al NMR study of

    alkali-activated slag. Cement and Concrete Research 33(5):

    769774.

    WHAT DO YOU THINK?

    To discuss this paper, please submit up to 500 words to

    the editor at www.editorialmanager.com/acr by 1 April2013. Your contribution will be forwarded to the

    author(s) for a reply and, if considered appropriate by

    the editorial panel, will be published as a discussion in a

    future issue of the journal.

    Advances in Cement Research

    Volume 25 Issue 1

    On a new hydraulic binder from stainless

    steel converter slag

    Pontikes, Kriskova, Cizer, Jones and Blanpain