interfacial reactions between coke/hdpe blends and ... - jst

7
41 © 2013 ISIJ ISIJ International, Vol. 53 (2013), No. 1, pp. 41–47 Interfacial Reactions between Coke/HDPE Blends and High Carbon Ferromanganese Slag Alexander BLAGUS,* James Ransford DANKWAH and Veena SAHAJWALLA Centre for Sustainable Materials Research and Technology (SMaRT@UNSW), School of Materials Science and Engineering, the University of New South Wales, UNSW Sydney, NSW 2052 Australia. (Received on May 22, 2012; accepted on September 24, 2012) The reduction of MnO in slag by blends of coke with high density polyethylene (HDPE) was investigated by the sessile drop method at 1 500°C in this study. The results show improved wettability and extents of reduction are realised with the use of an HDPE/coke blend in this system by comparison to reduction by pure coke, whereby increasing HDPE content resulted in further improvement in extent of reduction and increased wettability. The extensive devolatilisation from HDPE samples is the primary cause for these improvements, whereby the gasified HDPE created both CH4 and H2 reducing gases. Additionally, increased sample porosity allowed for improved wetting, and thus improved reduction capabilities. The dynamic contact angle between the carbon substrate and the slag varied, with HDPE samples ranging between 140°–60°, whilst the coke samples ranged between 160°–120°. The addition of HDPE allowed for the near complete reduction of MnO and partial reduction of SiO2 from the slag with distinct metallic regions of Mn–Si formed in the sample; regions containing pure Si were also found. KEY WORDS: manganese; MnO; carbothermal; reduction; metallurgical coke; waste plastics; polymer. 1. Introduction The use of waste polymeric materials within metallurgical industries is a relatively new concept, but the volumes of polymeric material consumption is continuing to increase, with almost 1.5 million tonnes of polymeric waste created annually in Australia alone (2010), of which only 16.2% is recycled. 1) In the field of electric arc furnace (EAF) steel- making, the utilisation of polymeric materials as a partial replacement for inject carbon (coke) has shown promise, with much research pertaining to the positive effects on reduction efficiencies. 2–8) These positive impacts should transfer through to other metallurgical fields, inciting research into parallel fields, including the production of high carbon ferromanganese (HC FeMn). HC–FeMn production worldwide is dominated by the submerged arc furnace (SAF) production route. 9) This pro- cess route is preferred over the blast furnace route for com- parative advantages such as reduced overall production costs, lower capital requirements, reduced capacities and lower ore and coke quality requirements. Despite the many advantages of this route, the SAF system for HC FeMn pro- duction is still carbon intensive, requiring approximately 300 kg of coke/tonne of hot metal; 10) this is a region wherein the introduction of polymeric materials may prove benefi- cial. Thermodynamically, carbon is required for the reduc- tion of MnO to either Mn metal or manganese carbides, the reactions following that shown in Eqs. (1), (2) and (3). The ΔG1 500°C values calculated for 1 500°C in Eqs. (1)–(10) clearly indicate the ability for these reactions to progress within the system at 1 500°C and were calculated through values presented in FACT-Sage 6.2 ((Thermfact/CRCT, Montreal, Canada, and GTT Technologies, Aachen, Germany) and experimental partial pressures (PCO, PCO 2 , PCH 4 ). PH 2 experimental values were not obtainable within the current experimental system. PH 2 values used were instead calculat- ed using the maximum volume of hydrogen attained through simultaneous devolatilisation of atomic hydrogen from an entire sample (Table 1). MnO (l) + C (s) = Mn (l) + CO (g) ................. (1) ΔG 1 500°C coke = –94 866 kJ/mol MnO ΔG 1 500°C B1 = –62 170 kJ/mol MnO ΔG 1 500°C B2 = –66 104 kJ/mol MnO ΔG 1 500°C B3 = –69 578 kJ/mol MnO 3MnO (l) + 4C(s) = Mn3C (l) + 3CO (g) ............ (2) ΔG1 500°C coke = –96 207 kJ/molMnO ΔG1 500°C B1 = –63 512 kJ/molMnO ΔG1 500°C B2 = –67 446 kJ/molMnO ΔG1 500°C B3 = –70 920 kJ/molMnO 7MnO (l) + 10C(s) = Mn7C3 (l) + 7CO (g) ........... (3) ΔG1 500°C coke = –103 412 kJ/molMnO ΔG1 500°C B1 = –70 717 kJ/molMnO ΔG1 500°C B2 = –74 651 kJ/molMnO ΔG1 500°C B3 = –78 124 kJ/molMnO The introduction of polymeric materials into the ferro- * Corresponding author: E-mail: [email protected] DOI: http://dx.doi.org/10.2355/isijinternational.53.41

Upload: others

Post on 09-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Interfacial Reactions between Coke/HDPE Blends and ... - JST

41 © 2013 ISIJ

ISIJ International, Vol. 53 (2013), No. 1, pp. 41–47

Interfacial Reactions between Coke/HDPE Blends and High Carbon Ferromanganese Slag

Alexander BLAGUS,* James Ransford DANKWAH and Veena SAHAJWALLA

Centre for Sustainable Materials Research and Technology (SMaRT@UNSW), School of Materials Science and Engineering,the University of New South Wales, UNSW Sydney, NSW 2052 Australia.

(Received on May 22, 2012; accepted on September 24, 2012)

The reduction of MnO in slag by blends of coke with high density polyethylene (HDPE) was investigatedby the sessile drop method at 1 500°C in this study. The results show improved wettability and extents ofreduction are realised with the use of an HDPE/coke blend in this system by comparison to reduction bypure coke, whereby increasing HDPE content resulted in further improvement in extent of reduction andincreased wettability. The extensive devolatilisation from HDPE samples is the primary cause for theseimprovements, whereby the gasified HDPE created both CH4 and H2 reducing gases. Additionally,increased sample porosity allowed for improved wetting, and thus improved reduction capabilities. Thedynamic contact angle between the carbon substrate and the slag varied, with HDPE samples rangingbetween 140°–60°, whilst the coke samples ranged between 160°–120°. The addition of HDPE allowed forthe near complete reduction of MnO and partial reduction of SiO2 from the slag with distinct metallicregions of Mn–Si formed in the sample; regions containing pure Si were also found.

KEY WORDS: manganese; MnO; carbothermal; reduction; metallurgical coke; waste plastics; polymer.

1. Introduction

The use of waste polymeric materials within metallurgicalindustries is a relatively new concept, but the volumes ofpolymeric material consumption is continuing to increase,with almost 1.5 million tonnes of polymeric waste createdannually in Australia alone (2010), of which only 16.2% isrecycled.1) In the field of electric arc furnace (EAF) steel-making, the utilisation of polymeric materials as a partialreplacement for inject carbon (coke) has shown promise,with much research pertaining to the positive effects onreduction efficiencies.2–8) These positive impacts shouldtransfer through to other metallurgical fields, incitingresearch into parallel fields, including the production of highcarbon ferromanganese (HC FeMn).

HC–FeMn production worldwide is dominated by thesubmerged arc furnace (SAF) production route.9) This pro-cess route is preferred over the blast furnace route for com-parative advantages such as reduced overall productioncosts, lower capital requirements, reduced capacities andlower ore and coke quality requirements. Despite the manyadvantages of this route, the SAF system for HC FeMn pro-duction is still carbon intensive, requiring approximately300 kg of coke/tonne of hot metal;10) this is a region whereinthe introduction of polymeric materials may prove benefi-cial. Thermodynamically, carbon is required for the reduc-tion of MnO to either Mn metal or manganese carbides, thereactions following that shown in Eqs. (1), (2) and (3). The

ΔG1 500°C values calculated for 1 500°C in Eqs. (1)–(10)clearly indicate the ability for these reactions to progresswithin the system at 1 500°C and were calculated throughvalues presented in FACT-Sage 6.2 ((Thermfact/CRCT,Montreal, Canada, and GTT Technologies, Aachen, Germany)and experimental partial pressures (PCO, PCO2, PCH4). PH2

experimental values were not obtainable within the currentexperimental system. PH2 values used were instead calculat-ed using the maximum volume of hydrogen attained throughsimultaneous devolatilisation of atomic hydrogen from anentire sample (Table 1).

MnO (l) + C(s) = Mn (l) + CO (g) ................. (1)

ΔG1 500°C coke = –94 866 kJ/molMnO

ΔG1 500°C B1 = –62 170 kJ/molMnO

ΔG1 500°C B2 = –66 104 kJ/molMnO

ΔG1 500°C B3 = –69 578 kJ/molMnO

3MnO (l) + 4C(s) = Mn3C (l) + 3CO (g) ............ (2)

ΔG1 500°C coke = –96 207 kJ/molMnO

ΔG1 500°C B1 = –63 512 kJ/molMnO

ΔG1 500°C B2 = –67 446 kJ/molMnO

ΔG1 500°C B3 = –70 920 kJ/molMnO

7MnO (l) + 10C(s) = Mn7C3 (l) + 7CO (g)........... (3)

ΔG1 500°C coke = –103 412 kJ/molMnO

ΔG1 500°C B1 = –70 717 kJ/molMnO

ΔG1 500°C B2 = –74 651 kJ/molMnO

ΔG1 500°C B3 = –78 124 kJ/molMnO

The introduction of polymeric materials into the ferro-* Corresponding author: E-mail: [email protected]: http://dx.doi.org/10.2355/isijinternational.53.41

Page 2: Interfacial Reactions between Coke/HDPE Blends and ... - JST

© 2013 ISIJ 42

ISIJ International, Vol. 53 (2013), No. 1

manganese system should allow the MnO reduction mech-anisms to utilise the high H/C ratio of the polymers. CH4

and H2 are the primary products from the pyrolysis of poly-meric materials, in addition to ethylene and other aromaticcompounds, as shown through a series of studies into thepyrolysis of high density polyethylene (HDPE) at 400–900°C.11–13) The reactions of MnO with methane shouldprogress as per the reactions presented by Eqs. (4), (5) and(6).14)

MnO (l) + CH4 (g) = Mn (l) + CO (g) + 2H2 (g) ....... (4)

ΔG1 500°C coke = –148 817 kJ/molMnO

ΔG1 500°C B1 = –131 772 kJ/molMnO

ΔG1 500°C B2 = –143 003 kJ/molMnO

ΔG1 500°C B3 = –143 481 kJ/molMnO

3MnO (l) + 4CH4 (g) = Mn3C (l) + 3CO (g) + 8H2 (g) ... (5)

ΔG1 500°C coke = –167 459 kJ/molMnO

ΔG1 500°C B1 = –155 631 kJ/molMnO

ΔG1 500°C B2 = –169 295 kJ/molMnO

ΔG1 500°C B3 = –168 147 kJ/molMnO

7MnO (l) + 10CH4 (g) = Mn7C3 (l) + 7CO (g) + 20H2 (g)

.......................................... (6)

ΔG1 500°C coke = –179 754 kJ/molMnO

ΔG1 500°C B1 = –169 414 kJ/molMnO

ΔG1 500°C B2 = –183 775 kJ/molMnO

ΔG1 500°C B3 = –182 969 kJ/molMnO

Anacleto et al.15) found that the presence of CH4 withMnO allowed for greater extents and rates of reduction incomparison to reduction under hydrogen or carbon monox-ide between 1 000–1 200°C. It was found that the higher car-bon activity of the methane allowed for improved rates ofreaction; they also found that the CH4 adsorbed onto theactive sites of the oxide surface, following which it decom-posed to adsorbed carbon and hydrogen gas. The study alsocommented that manganese oxide reduction was retardedwith excessive deposition of carbon onto the oxide surfacesfrom excessive CH4 decomposition.

The impact of hydrogen on the reduction of manganeseoxides has been studied in other works. Kononov et al.14)

showed that greater rates and extents in the reduction ofMnO were achievable with the presence of H2 in compari-son to reduction under the inert gases He and Ar between1 200–1 300°C. The suggested mechanism for the reductionof MnO by H2 involved the initial formation of methanethrough the reaction of hydrogen with graphite in the system(Eq. (7)), followed by the reaction of CH4 with MnO to formMn7C3 (Eq. (6)). The improved reaction rates and extents

were attributed to methane acting as a method of carbontransfer and the carbon activity in the methane containinggas phase approaching 1.0, allowing for manganese carbideformation.

C(s) + 2H2 (g) = CH4 (g) ........................ (7)

Recent work has also looked at MnO as a sessile dropand its interactions with a carbonaceous substrate. Safarianet al.16–18) used this method to show that MnO activity hadlittle impact on MnO-graphite reactions, the rate insteadbeing influenced by activation energy and the extent of slag-substrate surface area contact for temperatures between1 450, 1 500 and 1 600°C in Ar(g). Safarian et al.16) alsofound large wetting angles of slags with coke substrates(>150° for high MnO% slags), the wettability had an impacton the degree of surface area contact between the slag andthe solid carbonaceous material impacting on solid-liquidreactions such as that presented in Eqs. (1)–(3). Similarwork by Sun et al.19) also looked at the wetting of slags tocarbonaceous substrates (high MnO content slags), theyfound that with decreasing slag wetting angles (better wet-ting) the reduction rate increased for temperatures between1 450–1 550°C. Importantly they also found that the reductionof MnO, with regard to MnO reactivity, in the slag was a firstorder reaction with some of the rate control linked to themigration of MnO in the system towards reduced SiO2 (Si).

2. Experimental

2.1. Material CharacteristicsMetallurgical coke and three blends of coke-HDPE were

used as the carbonaceous material in this study; the proxi-mate and ultimate analyses of both the coke and HDPE arepresented in Table 1; the ash analysis of the coke sample ispresented in Table 2.

The coke sample was crushed in a tungsten carbide ringmill to a particle size of <75 μm, while the HDPE was sup-plied with particle sizes of <500 μm. These materials werethen blended in the proportions displayed in Fig. 1, addingproportionately larger quantities of HDPE resulted in sub-

Table 1. Chemical composition of Coke and HDPE.

Proximate and Ultimate analyses (%)

Material FixedCarbon

TotalCarbon

VolatileMatter Ash Sulphur Moisture Hydrogen

MetallurgicalCoke 73.2 77.4 6.1 17.2 0.36 3.4 1.13

HDPE 0.3 85.5 99.3 ~0 0.3 ~0 14.2

(-analysis conducted at Amdel Laboratories, Newcastle, Australia)

Table 2. Ash analysis of the metallurgical coke used.

Ash Analysis by XRF (%)

Components Metallurgical Coke

Silicon as SiO2 57.3

Iron as Fe2O3 5.4

Aluminium as Al2O3 26.5

Calcium as CaO 3.5

Magnesium as MgO 0.57

Sodium as Na2O 0.25

Potassium as K2O 0.60

Phosphorus as P2O5 0.57

Sulphur as SO3 2.78

Titanium as TiO2 1.12

Zinc as ZnO 0.03

(-analysis conducted at Amdel Laboratories, Newcastle,Australia)

Page 3: Interfacial Reactions between Coke/HDPE Blends and ... - JST

ISIJ International, Vol. 53 (2013), No. 1

43 © 2013 ISIJ

stantial substrate structure collapse following reaction, thuslimiting polymer addition for this experimental setup. Theblended carbonaceous materials were mixed in a rollingmill for greater than 4 hours, following which substrates ofthe mixture were created. The substrates contained 1.6 g ofcarbonaceous material which was uniaxially pressed at 60kN/cm2 into cylindrical discs of 20 mm in diameter.

A synthetic slag was used in this study, with the compo-sition similar to industrial HC FeMn slags and syntheticslags presented in literature [Refs. 10), 19), 22)]. The slagwas prepared by melting the oxide constituents at 1 600°Cin air within a muffle furnace, following which the slag wasremoved and quenched in water. This process was repeateda further 2 times and following the final quench the samplewas powdered in a ring mill. The composition of the slagwas assessed following preparation and the composition ispresented in Table 3 along with other examples of slags pre-sented in previous studies.

2.2. Carbon/Slag InteractionsThe sessile drop technique was utilised for investigating

the interactions between the slag and a substrate of carbo-naceous materials at 1 500°C, similar to the approach byother investigators.16–19) The study involved the use of a lab-oratory scale horizontal alumina tube furnace (schematicsshown in Fig. 2). The sample assembly was composed of agraphite rod with a tray that was filled with high purity alu-mina powder on which the substrate with slag was placed.The assembly and sample were held in the cold zone at~300°C with an argon gas flow of 1 L/min until the atmo-sphere was purged and the sample had heated somewhat forover 15 minutes, the assembly was then inserted into the hotzone at 1 500°C where reactions continued for 20 minutes.Following reaction the sample was removed and held in thecold zone for over 15 minutes so as to avoid re-oxidation ofthe reduced sample. The system was continuously purged

with argon, allowing for low O2, CO and CO2 partial pres-sures.

A quartz window at the rear of the furnace allowed for insitu imaging of the samples; images were obtained frequent-ly (20 second intervals) during reaction and these were usedto assess the wetting angle changes between the slag and thedifferent substrates. The final product was further analysedby scanning electron microscopy (SEM – Hitachi 3400using a Bruker EDS system) whereby two viewpoints of thesamples were taken. Topographic analysis of the slag/sub-strate involved setting the samples partially in resin leavingthe sample surface unchanged, providing insights into theinteraction of the slag with the surface of the substrate.Cross-sectional analysis required the sample to be fully setin resin and cross sectioned laterally to analyse the interfa-cial region of the slag with the carbonaceous substrates. AnIR gas analyser (Advance Optima model ABB® AO2020)attached to this system was used to monitor CH4, CO andCO2 gases produced by the reactions and the results wererecorded in a data-logging computer.

3. Results and Discussion

3.1. Infrared (IR) Gas AnalysisFigure 3 shows various plots of the concentration of CO,

CO2 and CH4 as a function of time during the reduction ofSiMn slag by each carbonaceous reductant. As can beobserved in Fig. 3(a), CO2 is the dominant component of theoff-gas from the reduction of the slag by coke, most likelythrough a reaction of the form

MnxOy + CO = MnxOy–1 + CO2................. (8)

Minor amounts of CH4 were also detected from the reduc-

Fig. 1. Relative proportions of coke and HDPE present in the dif-ferent blends used for this study.

Table 3. Slag analyses of relevant industrial and synthetic slag samples.

Slag analysis by XRF (%)

TEMCO(Australia)10)

Eramet(Norway)22)

IndependentSynthetic Slag18)

Present StudySynthetic Slag*

MnO 34.9 45.6 45 39.4

SiO2 25.9 22.9 27.5 27.2

Al2O3 15.1 10.5 11 12

Fe2O3 0 0.07 0 0.2

MgO 0.9 4.6 0 2.0

CaO 14.6 13.9 16.5 15.9

Total 91.4 97.6 100 96.7

(* -analysis conducted at Analytical Centre, UNSW-Sydney, Australia)

Fig. 2. Horizontal tube furnace schematic diagram.

Page 4: Interfacial Reactions between Coke/HDPE Blends and ... - JST

© 2013 ISIJ 44

ISIJ International, Vol. 53 (2013), No. 1

tion reaction by coke probably due to devolatilisation (coke:6.1% volatile matter + 1.13% H). Figs. 3(a)–3(c) illustratesthe gas generation behaviour when the slag is reduced byBlends 1–3, respectively. Unlike in the case of reduction bycoke, CO2 is no longer the dominant component as onlyminor amounts are detected in the off-gas. The significantdecrease in CO2 emission coupled with the massive amountsof both CO and CH4 is an indication that the polymer under-goes thermal decomposition to liberate significant amountsof CH4, as was observed by Nishioka et al.20) and Dankwahet al.21) At this stage it is valid to speculate that part of theextra CO observed in the reduction by the blends could beattributed to reduction of MnxOy by CH4, as shown in Eq. (9).

MnxOy + CH4 = MnxOy–1 + CO + 2H2 + ................... (9)

Gas generation from the reduction of the slag by theblends appears to be more significant than from the reduc-tion of the slag by coke. An improved rate of reductionshould therefore be expected when coke is blended with thepolymer.

3.2. Interfacial Reactions in the Slag/Carbon SubstrateSystem

3.2.1. Wetting BehaviourThe substrates that contained polymeric materials exhib-

ited increased wetting by comparison to slags reacting onpure coke substrates, as shown through both Figs. 4 and 5.

Fig. 3. Concentration (vol%) of CO, CO2 and CH4 in the off-gas as a function of time during the reduction of the syntheticslag by each reductant at 1 500°C.

Fig. 4. Images depicting the change in wetting behaviour with time of the liquid slag on different substrates (coke, Blend1, Blend 2 and Blend 3).

Fig. 5. Variation of wetting angles with time for coke and Blends1, 2 and 3.

Page 5: Interfacial Reactions between Coke/HDPE Blends and ... - JST

ISIJ International, Vol. 53 (2013), No. 1

45 © 2013 ISIJ

Whilst increasing reaction times resulted in improved wet-ting by the slag, the wetting angle of B3 displayed the bestwetting improvement with reaction time, followed by B2,B1 and then coke. The variation between the wetting anglesof coke compared to B1, B2 and B3 are fairly clear, wherebyat 15 minutes the slag is very wet to the substrate in B1, B2and B3, and following 20 minutes of reaction time the slagsin B2 and B3 have penetrated into the coke-polymer blendsubstrate allowing for a more intimate contact between slagand the carbon components of the substrate, allowing forimproved reduction and therefore improved gas generationrates. The variation in wetting angles on the pure coke sub-strates did not change as fast as those of B1, B2 and B3. Thecoke samples are shown to compare well to previous workby Sun et al.19) (Fig. 6).

The improved wetting primarily comes about through theincreased porosity of the samples, resultant from the devola-tilisation of polymeric materials in B1, B2 and B3. Examplesof the increased porosity are shown in Fig. 7, where visualsurface macro porosity was measured at 1.0% for coke, 2.4%for B2 and 3.6% for B3. The pore network also extendsbelow the surface as shown through the SEM images pre-sented in Fig. 8. The increase in porosity has the chance toimpact on wettability with increased surface roughness,inducing improved wetting and thus improving reactionrates and extents. This noted, similar work by Safarian et al.17)

found that more than the surface roughness or CO2 reactiv-ity of a carbonaceous substrate, the ‘type’ of carbon had alarger effect, suggesting that the porosity increase may onlypartially impact on wetting and/or reduction of high MnOcontaining slags.

3.2.2. Topical Slag–Substrate InteractionsThe topographic images of the reaction zone give a clear

indication of the interactions between the metal, slag and

carbon components; shown in Fig. 9 are examples of coke,B1, B2 and B3. Of all the samples, coke showed the leastreduction, with the surface droplet shown in Fig. 9(d) com-prised primarily of unreduced slag, with a small amount ofcarbonaceous material present on its surface. By compari-son, the HDPE containing blends, B1, B2 and B3, showedvastly superior degrees of wetting and reduction, presentedthrough the clearly defined metallic droplets formed on thesubstrate surface. EDS analysis confirmed the presence ofnon-oxidised Mn and Si in the metallic droplets and a dis-tinct lack of other slag components. The presence of a car-bon region on top of some metallic droplets in the threeHDPE containing blends, suggests the formation of manga-nese carbides in the system. This noted; carbon content isdifficult to clearly ascertain as EDS is not very effective forthe determination of the presence of lower atomic numberelements, such as carbon.

Silicon metal was also discernible through EDS analysis,observed primarily at the edges of metallic regions. Anexample of the Si crystals is shown with the B3 sample inFig. 9(c). SiO2 will also be reduced by carbon at 1 500°Cwhilst MnO reduces, following Eq. (10). Following the for-mation of Si metal, the Si reduces the remaining MnOthrough Eq. (11). When the MnO is fully consumed in thereaction system, any further SiO2 reduction produces siliconmetal to be retained in the metal phase, either as Si metal or

Fig. 7. Surface variation with HDPE content a) coke, b) Blend 1and c) Blend 3. All samples are 20 mm in diameter.

Fig. 6. Variation of contact angle with time (First 20 min of reaction).

Fig. 8. SEM images displaying interior porosity variation with HDPE content: a) coke, b) Blend 1 and c) Blend 3.

Page 6: Interfacial Reactions between Coke/HDPE Blends and ... - JST

© 2013 ISIJ 46

ISIJ International, Vol. 53 (2013), No. 1

as a silicon carbide, thus Si metal presence suggests thecomplete reduction of MnO to Mn or its carbides.

SiO2+2C=Si+2CO ......................... (10)

ΔG1 500°C coke = –79 166 kJ/molMnO

ΔG1 500°C B1 = –13 776 kJ/molMnO

ΔG1 500°C B2 = –21 643 kJ/molMnO

ΔG1 500°C B3 = –28 590 kJ/molMnO

Si+ MnO=Mn+ SiO........................ (11)

3.2.3. Interfacial Interactions between Slag, Metal andSubstrate

Figure 10 displays the cross-sections of the carbon/slag/metal interface for B1, B2 and B3. The coke samples dis-played coke minimal wetting with few to no metallicregions, B1, B2 and B3 show extensive wetting beyond theoriginal droplet and also a high degree of slag penetrationinto the substrate. The increased penetration allows forgreater contact between the slag and the carbonaceous sub-strate, providing improved carbon-slag reactions, such asthose described in Eqs. (1), (2), (3), and (10). Indeed Sun etal.19) similarly found a distinct relationship where largerwetting angles allowed for increased rates of reduction. The

reactions would improve with bettered contact between theMnO in the slag and its source of carbon, the effect of theslag penetration is shown by the degree of metal nucleationpresented in the HDPE containing samples shown in Fig.10, with most nucleation sites centring around sources ofsolid carbon (coke) and having prills sized between 150–450 nm throughout the slag pool.

The reaction rates of Eqs. (1), (2), (3), and (10) shouldalso increase with improved slag/carbon interaction, thisgiving greater opportunities for carbon to dissolve into themetal phase. This is shown through the distribution of car-bonaceous material within the slag pools; a line has beendrawn on the images in Fig. 10 indicating the region whereno carbonaceous material resides within the slag pool. Inthese regions the slag/metal dissolved any carbon presentand allowed for Eqs. (1), (2), (3), and (10) to proceed moreeffectively. The outer extremities of the slag pool (outsidethe line) show regions where the slag/metal has yet to fullydissolve/react with the carbon materials present.

Also noticeable from the cross-sectional images of B1,B2 and B3 is the formation of zones within the metalregions containing higher concentrations of Si. Presentedbelow in Figs. 11 and 12 is the line scan of a metallic regiontypical of Blends 1–3. The darker zones within the metal

Fig. 9. Topographical views of the metal/slag region for substrates (clockwise from top left) a) Blend 1, b) Blend 2, c)Blend 3 and d) coke.

Fig. 10. BSE images of the interfacial region between the slag pool and the carbonaceous substrates for a) Blend 1, b)Blend 2 and c) Blend 3.

Page 7: Interfacial Reactions between Coke/HDPE Blends and ... - JST

ISIJ International, Vol. 53 (2013), No. 1

47 © 2013 ISIJ

region contain higher concentrations of Si, portentous ofSiO2 reducing to Si following the complete reduction ofMnO; thus these zones containing greater Si concentrationssuggest that most MnO has been reduced from the system.Additionally, EDS analysis of the slag around the metallicregion suggested minimal, if any, manganese presence.Finally, the EDS analysis revealed a slight increase in thecarbon readings within the metal region, whilst EDS is notvery accurate for such reading this does suggest that carbideformation occurred.

4. Conclusions

The introduction of polymeric materials along with cokeprovided for improved wetting of the carbonaceous sub-strate by the slag and allowed for a greater extent of reduc-tion of MnO and SiO2 from the slag as shown through IRgas analysis with the greatest volume of CO being releasedfrom polymer blended samples.

Improved wetting by slag was found to occur when poly-

meric materials were blended with coke to form the carbo-naceous substrate by comparison to pure coke substrates.The best wetting occurred in the Blend 3 substrate, contain-ing the largest proportion of HDPE followed by Blend 2,Blend 1 and then coke. The improved wetting with HDPEaddition resulted from a mixture of increased porosity anda change in the characteristics of the carbonaceous substrate.

The characteristics of the sample following reactionshowed near total reduction of MnO from the slag, with par-tial reduction of SiO2 to Si.

AcknowledgementsFinancial support for this project, provided by OneSteel,

is greatly appreciated. The authors appreciatively acknowl-edge access to, and technical support from, the UNSW nodeof the Australian Microscopy & Microanalysis ResearchFacility (AMMRF), UNSW.

REFERENCES

1) 2010 National Plastics Survey: July 2009 to June 2010 surveyperiod - Final Report, Plastics and Chemicals Industries Association(PACIA) Inc., Melbourne, Australia, (2010).

2) V. Sahajwalla, L. Hong and C. N. Saha: AISTech, AIST, Associationfor Iron and Steel Technology, Warrendale, PA, USA, (2005), 597.

3) V. Sahajwalla, R. Khanna, M. Rahman, Z. Huang, E. Tanaka, N.Saha-Chaudhury, D. Knights and P. O’kane: AISTech, Associationfor Iron and Steel Technology, Warrendale, PA, USA, (2006), 547.

4) M. Rahman, V. Sahajwalla, R. Khanna, N. Saha-Chaudhury, D.Knights and P. O’kane: AISTech, Association for Iron and SteelTechnology, Warredale, PA, USA, (2006), 491.

5) S. Gupta, V. Sahajwalla and J. Wood: Energ. Fuel., 20 (2006), 2557.6) M. Zaharia, V. Sahajwalla, B.-C. Kim, R. Khanna, N. Saha-Chaudhury,

P. O’kane, J. Dicker, C. Skidmore and D. Knights: Energ. Fuel., 23(2009), 2467.

7) M. Zaharia, V. Sahajwalla, R. Khanna, P. Koshy and P. O’kane: ISIJInt., 49 (2009), 1513.

8) V. Sahajwalla, M. Zaharia, S. Kongkarat, R. Khanna, N. Saha-Chaudhury and P. O’kane: Energ. Fuel., 24 (2010), 379.

9) L. A. Corathers: U. S. Geological Survey, (2008), 17.10) J. Bourke, S. Ganguly and D. Retallick: AusIMM, 2 (1993), 1223.11) F. J. Mastral, C. Esperanza, C. Berrueco, M. Joste and J. Ceamanos:

J. Anal. Appl. Pyrol., 70 (2003), 1.12) L. Martinez, A. Aguado, A. Moral and R. Irusta: Fuel Process.

Technol., 92 (2011), 221.13) A. Marcilla, M. D. R. Hernandez and A. N. Garcia: J. Anal. Appl.

Pyrol., 79 (2007), 424.14) R. Kononov, O. Ostrovski and S. Ganguly: Metall. Mater. Trans. B,

39 (2008), 662.15) N. Anacleto, O. Ostrovski and S. Ganguly: ISIJ Int., 44 (2004), 1480.16) J. Safarian and L. Kolbeinsen: ISIJ Int., 48 (2008), 395.17) J. Safarian, L. Kolbeinsen, M. Tangstad and G. Tranell: Metall.

Mater. Trans. B, 40 (2009), 929.18) J. Safarian, G. Tranell, L. Kolbeinsen, M. Tangstad, S. Gaal and J.

Kaczorowski: Metall. Mater. Trans. B, 39 (2008), 702.19) H. Sun, M. Y. Lone, S. Ganguly and O. Ostrovski: ISIJ Int., 50

(2010), 639.20) K. Nishioka, T. Taniguchi, Y. Ueki, K.-I. Ohno, T. Maeda and M.

Shimizu: ISIJ Int., 47 (2007), 602.21) J. R. Dankwah, P. Koshy, N. M. Saha-Chaudhury, P. O’kane, C.

Skidmore, D. Knights and V. Sahajwalla: ISIJ Int., 51 (2011), 498.22) T. A. Skjervheim and S. E. Olsen: INFACON 7, SINTEF, Trondheim,

Norway, (1995), 631.

Fig. 11. SEM back-scattered electron image from Blend 2. Thedarker zones within the metal region correlate with zonescomparatively higher in Si, as presented in Fig. 12.

Fig. 12. EDS analysis for C, Si, Mn and O along the line from Fig. 11.