Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter February 28, 2015

Reaction between Steel-Making Slag and Carbonaceous Materials While Mixing with High Density Polyethylene

  • Lan Hong and Veena Sahajwalla EMAIL logo

Abstract

Since the beginning of the extensive applications in numerous high temperature processes such as iron- and steel-making, coke-making etc. partly in the place of coke, the investigation into the reaction mechanism of waste plastics has become increasingly necessary. In this paper a fundamental study on the behavior of a typical component of waste plastics, high density polyethylene (HDPE), in a mixture with coke at a 1:1 ratio in mass base was conducted during the reaction with iron oxide in steel-making slag at 1823 K and was compared with coke and graphite. The reaction mechanism of carbonaceous materials was analyzed based on the contents of CO and CO2 in the off-gas monitored by an infrared (IR) gas analyzer. It is clear from the results that the reaction of HDPE and coke mixture with steel-making slag approached equilibrium of the Boudouard reaction more quickly and closely than coke or graphite.

Introduction

Carbonaceous materials are widely utilized in iron-making, steel-making and power generation processes, etc. However, with the ever-increasingly stringent environmental regulations about green house gas CO2 emission, the utilization of carbon must be cut. Some candidates to replace carbon can be named as waste plastics, in which element hydrogen can function similarly as element carbon so that less CO2 emission can be attained, and renewable bio-mass, which releases no CO2 overall.

On the other hand, value-added reutilization of waste plastics has attracted tremendous global attention. Waste plastics have been studied for applications in iron- and steel-making processes [1] such as injection into blast furnaces [2] and electric arc furnaces [3], as a partial replacement of coke, as well as in the coke-making industry [4, 5], and in manufacturing advanced carbon nano-structures [6] (cited as carbon nano-tubes or graphenes). Although it is mainly considered a pollutant commonly today, the potential of waste plastics as a strategic national resource has been considered in many nations. This is seen in the increasing focus into the search for ways to utilize waste plastics more efficiently.

A most recent trend of waste plastic applications is to be utilized as a reducing agent since the main elements in plastics, carbon and hydrogen, are extensively used to reduce oxides. It has been found that waste plastic can reduce iron oxide from lower temperature than graphite or coke [7], and the partial replacement of coke by PE in EAF practice can stabilize the operation by introducing controllable slag foaming originated from the gas release from the reaction between PE/coke and iron oxide in molten slag [3]. A comprehensive understanding about the reaction behavior of waste plastics at high temperatures is one of the key issues in assisting in optimizing efficient and economical application of waste plastics in steel-making processes.

HDPE is a typical component of post-consumer plastics. Therefore, the study about the reaction behavior of HDPE definitely contributes to the knowledge of waste plastic applications. Previously the current authors have reported the successful injection of HDPE to EAF steel-making practices. In this work the interaction between HDPE and iron oxide in steel-making slag was studied at high temperature of 1823 K for a more comprehensive understanding about the process. A comparison of the reaction behavior of HDPE and coke mixture with two commonly used carbonaceous materials, graphite and coke, was conducted as well. HDPE was added in a mixture with coke to avoid clogging, which usually happens when only HDPE is injected through pipes in industrial practices.

Experimental

Experimental apparatus

A horizontal resistance furnace was employed in the current study to provide the high temperature environment for the reaction, of which schematic diagram is shown in Figure 1. Both ends of the furnace tube were sealed to ensure the furnace was air-tight, allowing the easy manipulation of the internal environment. An alumina sample holder, connected to a stainless steel rod, was used to transport samples to and from the isothermal zone of the furnace. The outlet gas was led to a CO/CO2 IR gas analyzer to monitor the off-gas composition for further analyzing the reaction mechanism.

Figure 1 Schematic diagram of experimental setup
Figure 1

Schematic diagram of experimental setup

Experimental materials

Synthesized graphite powder, industry coke fines and HDPE particles were chosen as the starting raw materials for the study. The purity of graphite (mean size 44 µm) and HDPE (mean size 3 mm) was more than 99.9%. The composition of the coke is listed in Table 1. The coke was grinded to get an average size of about 100 µm before use. Three types of carbonaceous blocks around 10 g each were prepared for the experiments using high pressure hydraulic pressing (2.2 ╳ 108 Pa). These include the blocks of graphite, coke and the mixture of HDPE and coke with a mixing ratio of 1:1 based on the results of a previous work that it is the maximum mixing ratio for HDPE without causing operation disorders. The mixture of HDPE and coke was made by mixing the pre-weighed grinded coke powder and HDPE homogeneously before being hot-pressed at 423 K. Steel-making slag composed of 29.5% CaO, 34.7% FeO, 8.2% Al2O3, 9.3% MgO, 12.9% SiO2 and 5.4% MnO was used to react with carbonaceous materials. One thing should be mentioned that the iron oxide is present in the steel-making slag as both FeO and Fe2O3. Due to the difficulty distinguishing these two phases, iron ions are counted only as FeO. The initial weight of slag was around 0.1 g. The experiments were carried out in argon atmosphere and the gas flow rate was set at 1.67 ╳ 10−5 m3/s, equivalent to a linear velocity of 8.5 ╳ 10−3 m/s in a Φ50 ╳ 1000 mm corundum tube.

Table 1

Compositions of industry coke (%mass)

Fixed carbon83Ash composition: 50.1 SiO2, 38.2
Ash15Al2O3, 5.3 Fe2O3, 2.0 CaO, 1.7 TiO2,
Volatile11.5 P2O5, 0.4 MgO, 0.3 Na2O, 0.4
Moisture1K2O, 0.1 SO3

Experimental procedure

After around 0.1 g of slag was put on a pre-weighed carbonaceous block located on the alumina sample holder in the furnace, the furnace tube was sealed. Then the sample was pulled back to the cold zone close to the water-cooled front flange to keep it at low temperature before the commencement of the experiment. Then Ar gas was purged to remove the air inside the furnace tube for 1800 s before heating was started. When the temperature of the isothermal zone reached the pre-set temperature of 1823 K, the sample was pushed to the middle of the furnace tube to initiate the experiment, and at this moment the CO/CO2 IR gas analyzer was started as well to monitor the CO and CO2 contents in the off-gas originated from the reaction between carbon and iron oxide in the slag. After a certain period of time, the reaction was stopped by pulling the sample holder, on which the sample was located, back to the cold zone.

Experimental results and discussion

The measured CO and CO2 in the off-gas are plotted against reaction time for the carbonaceous substrates of HDPE and coke mixture together with coke and graphite in Figure 2. It is assumed that the measured gas composition was of the local source since the atmosphere gas flow rate was set large to eliminate the influence of gas phase diffusion on the overall reaction rate. As similar initial weight for all the carbonaceous substrates was used, it is reasonable to compare the reaction rate qualitatively by the CO and CO2 contents in the off-gas shown in the graphs for different carbonaceous materials.

Figure 2 (a) Change of CO and CO2 contents in the off-gas with time for graphite. (b) Change of CO and CO2 contents in the off-gas with time for coke. (c) Change of CO and CO2 contents in the off-gas with time for the mixture of HDPE and coke
Figure 2 (a) Change of CO and CO2 contents in the off-gas with time for graphite. (b) Change of CO and CO2 contents in the off-gas with time for coke. (c) Change of CO and CO2 contents in the off-gas with time for the mixture of HDPE and coke
Figure 2 (a) Change of CO and CO2 contents in the off-gas with time for graphite. (b) Change of CO and CO2 contents in the off-gas with time for coke. (c) Change of CO and CO2 contents in the off-gas with time for the mixture of HDPE and coke
Figure 2

(a) Change of CO and CO2 contents in the off-gas with time for graphite. (b) Change of CO and CO2 contents in the off-gas with time for coke. (c) Change of CO and CO2 contents in the off-gas with time for the mixture of HDPE and coke

There are two distinguishable reaction patterns classified according to the dominant reaction products, which are called the coke type and the graphite type in the present study. The output of the oxidation reaction of graphite was mainly CO2 in a relatively long period of the reaction time, whereas the dominant reaction product for the oxidation reaction of the mixture of HDPE and coke or coke only was CO almost throughout the whole duration of the reaction except a few seconds in the very beginning. The amounts of CO and CO2 originating from the reaction between the slag and coke containing substrates, HDPE and coke mixture or coke only, were much greater than those from the reaction between the slag and graphite. Both CO and CO2 contents rose to their respective peaks quickly after the start of the reaction and fell down steeply to a gradually changing stage in the case of graphite substrate. In contrast to this, the CO content in the off-gas in the cases of HDPE and coke mixture or coke only rose to a near plateau maximum value zone lasting for around 600 s, then from the maximum value zone dropped to a gradually decreasing zone in a very short time; on the other hand, CO2 content increased to a peak value after starting the reaction, and then decreased gradually along the reaction time. The CO and CO2 contents in the off-gas for the substrate of HDPE and coke mixture, and coke only substrate follow a similar trend.

The reactions between carbonaceous materials and iron oxide (Fe2O3 and FeO) in molten slag are generally accepted following the four reactions forming CO and CO2 as indicated below step-wisely, which is Fe2O3 is reduced firstly to FeO and then to Fe.

(1)[8]C(s)+Fe2O3(l)=2FeO(l)+CO(g)ΔG0=187610229.07TJ/mol
(2)[8]Cs+2Fe2O3l=4FeOl+CO2gΔG0=208656287.15TJ/mol
(3)[8]C(s)+FeO(l)=Fe(l)+CO(g)ΔG0=180226139.45TJ/mol
(4)[8]C(s)+2FeO(l)=2Fe(l)+CO2(g)ΔG0=10072954.22TJ/mol

Since the CO and CO2 formation reactions (5) and (6) have been proved to be primary ones by isotopic studies [9], the above four reactions (1) to (4) should occur independently.

(5)[8]C(s)+12O2(g)=CO(g)ΔG0=11439185.77TJ/mol
(6)[8]C(s)+O2(g)=CO2(g)ΔG0=3953460.54TJ/mol

However, the following three reactions may occur but not necessarily depending on the conditions.

(7)[8]CO(g)+Fe2O3(l)=2FeO(l)+CO2(g)ΔG0=2104658.07TJ/mol
(8)[8]CO(g)+FeO(l)=Fe(l)+CO2(g)ΔG0=24894+31.55TJ/mol
(9)[8]C(s)+CO2(g)=2CO(g)ΔG0=166565171.00TJ/mol

The mean size of graphite powder used in this work is 44 µm and the average grain size of coke is 100 µm, the specific surface area of coke is less than that of graphite. However, coke containing substrates, coke and the mixture of HDPE and coke, released CO and CO2 more quickly than graphite substrate. This is because graphite is chemically more stable than coke. The differences in the crystalline structure of graphite and coke are considered to be one of the causes. There are two types of planes in crystallized or partially crystallized carbonaceous materials: basal planes and prismatic or edged planes. The surface energies of these two types of crystallographic direction differ largely: about 0.11 J/m2 in the basal planes and 5 J/m2 in the prismatic planes [10]. Therefore basal planes are very stable and prismatic or edged planes tend to have lower reaction activation energies. The crystalline size of graphite is larger than that of coke, which also has a highly scattered crystalline direction, as shown in Figure 3. For the typical carbonaceous materials peak at the scanning angle of around 26°, it is very sharp for graphite suggesting the crystalline size of graphite is very large and the crystals are well-oriented; while the peak of coke is very broad, corresponding to the fact that in coke only a small portion is graphitized whilst its dominant part remains amorphous with small crystals. As a result, coke is more reactive than graphite due to its larger proportion of edged planes and hence more products CO and CO2 are generated in the case of coke compared to graphite. The reaction of prismatic or edged planes may be predominant during the quick reaction period, while the slow reaction stage may be attributed to the fact of the reaction area being occupied mostly by basal planes after the prismatic or edged planes initially exposed to the atmosphere have been consumed. Another reason for the quicker reaction rate of coke than graphite may exist in the fact that coke contains 15% ash composed of oxides such as lime and iron oxide which may act as catalysts during the reaction [11].

Figure 3 XRD results of various carbonaceous materials
Figure 3

XRD results of various carbonaceous materials

The standard free energy changes of CO and CO2 generation reactions (5) and (6) are plotted in Figure 4. It is clear from this figure that the formation reaction of CO2 proceeds preferably compared to the formation reaction of CO as expressed by eq. (5). Thus, more CO2 was generated than CO in the initial stage of the oxidation for all the carbonaceous substrates studied in the current research. This should correspond to the reduction period of Fe2O3 to FeO according to eq. (7). However, the duration of favored production of CO2 over CO was longer for graphite whereas in the cases of coke and the mixture of HDPE and coke this phenomenon only appeared for seconds in the very beginning of the reaction. This may be due to the fact that graphite is less active than coke, therefore longer time is needed to reduce all Fe2O3 to FeO for graphite than coke. Afterwards CO production dominated over CO2 production, especially for coke and the mixture of HDPE and coke, where CO level was markedly higher than CO2.

Figure 4 Standard free energy change as a function of temperature
Figure 4

Standard free energy change as a function of temperature

The index of Pco/Pco2, an indicator for the ratio of CO to CO2 partial pressure, is plotted as a function of time in Figure 5. The reaction time at which Pco/Pco2 = 1 for all of these three carbonaceous substrates shows no correlation with their respective reaction time in Figure 2(a)–(c), at which the oxidation reaction transits from the fast reaction rate zone to the slow reaction rate zone of the dominant oxidation product (CO2 for HDPE + coke mixture and coke, CO for graphite).

Figure 5 Change of Pco/Pco2 value with time
Figure 5

Change of Pco/Pco2 value with time

Since carbon was in excess near the reaction sites, the reactions (7) and (8) may not necessarily proceed while the Boudouard Reaction or the solution-loss reaction (9) occurs inevitably. Therefore the solution-loss reaction of carbon is considered to be the reason of CO overtaking CO2 later on.

The solution-loss reaction (9) plays a key role in many carbothermic reduction processes of oxides and is judged as at least one of the main rate controlling steps by many researchers [12]. The value of pco2/pco2 for the reaction system of HDPE + coke mixture, coke and graphite with iron oxide in molten slag is calculated and plotted in Figure 6. In this graph, the dotted line is the value of pco2/pco2 when the solution-loss reaction is at equilibrium. The mixture of HDPE and coke approached the equilibrium of the solution-loss reaction most quickly from the beginning followed by coke, while the oxidation of graphite hardly reached the equilibrium of the solution-loss reaction. As a result, CO content soon surpassed CO2 content in the gas phase in the cases of HDPE + coke mixture and coke, whereas for graphite and iron oxide reaction system more CO2 was detected than CO in the off-gas during most time of the reaction. The fact that HDPE + coke mixture reached more quickly than coke is correlated to the pyrolysis characteristics of HDPE at high temperature. Many researchers have investigated the pyrolysis behavior of PE. I. Kayacan et al. [13] showed that the pyrolysis of HDPE proceeded vitally from below 700 K. At low temperatures, the pyrolysis products included solid carbon residue, oil or wax and light organic and inorganic gases [14] and their proportions varied with experimental conditions such as temperature, heating rate, presence of catalysts, etc. With the increase in the temperature, the wax or oil started to break down to small molecules. When the temperature was very high as iron- and steel-making ones, the reaction between PE and hematite yielded H2 and CO with small amounts of H2O and methane in the gas phase as shown by T. Matsuda et al. [15]. H2O and CO gas must be the reaction product of hydrogen-iron oxide and carbon-iron oxide, respectively. For the current case, since the temperature is very high as 1823 K, it should not bring much dispute if reaction (10) is used to represent the overa(CH2)n=nCs+nH2g ll pyrolysis process of HDPE to simplify the role played by HDPE in this study as to provide reducing agents H2 and solid residue carbon.

(10)(CH2)n=nCs+nH2g
Figure 6 Comparison of pco2/pco2$${\rm{p}}_{{\rm{co}}}^{\rm{2}}/{{\rm{p}}_{{\rm{c}}{{\rm{o}}_2}}}$$ of carbonaceous substrates with the equilibrium value of solution-loss reaction
Figure 6

Comparison of pco2/pco2 of carbonaceous substrates with the equilibrium value of solution-loss reaction

Although the current work has not monitored the H2 and H2O contents in the off-gas, the release of H2 gas from PE decomposition has been confirmed by many other researchers. T. Murakami et al. [7] indicated that PE played a significant role for the reduction of iron oxide by supplying reducing gases, H2 and CO, through pyrolysis when utilizing PE in combination with graphite as reducing agents. The temperature of the gas generation peak originated by the reduction of graphite, 1370 K, was much higher than those of the other two gas generation peaks, 770 K and 1070 K, caused by the reaction of PE. They also showed the residual carbon after PE pyrolysis was ready to be oxidized by CO2 from a much lower temperature than coke or graphite, which implies that the derived carbon from PE pyrolysis is chemically more active than coke and graphite. In addition, A. Bazargan et al. [6] reviewed that the residual carbon after PE pyrolysis was very porous and possessed a large specific surface area compared to coke or graphite. The current result graphed in Figure 6 that the mixture of HDPE and coke approached the equilibrium of solution-loss reaction the most quickly among the carbonaceous materials investigated in this study can be explained correspondingly.

Another important result to be pointed out in Figure 6 is that once in a short time period, the mixture of HDPE and coke showed a larger pco2/pco2 value than the equilibrium value of the solution-loss reaction. This may be due to the fact that hydrogen, one of the pyrolysis products of HDPE according to reaction (10), can react further with CO2 according to eq. (11) to give a higher CO content and a lower CO2 content compared to their respective equilibrium values according to the solution-loss reaction.

(11)CO2g+H2g=COg+H2Og

The release of hydrogen gas from HDPE is a very fast process starting from low temperatures as low as 650 K [16]. Therefore, it is highly possible that after 1000 s, there were no more hydrogen in the reaction system. However, the pco2/pco2 value of the mixture of HDPE and coke was still above that of coke, much closer to the equilibrium value of the solution-loss reaction than that of coke oxidation at the latter stage of the reaction. This is probably due to that in the mixture of HDPE and coke, after the pyrolysis of HDPE, there was still porous solid carbon residue chemically more active than coke or graphite getting involved in the reaction in the system.

Conclusions

The reaction behavior of HDPE in the mixture with coke was compared to graphite and coke by oxidizing these carbonaceous materials by iron oxide in molten slag at high temperature of 1823 K. CO was the main gas product in both cases of HDPE + coke mixture and coke, whereas CO2 was released dominantly when graphite reacted with iron oxide. As a reducing agent, HDPE showed greater oxygen affinity than both coke and graphite probably due to the fact that the main pyrolysis products of HDPE, both porous solid carbon residue and hydrogen gas, were chemically more active than coke and graphite so that the solution-loss reaction was accelerated consequently.

References

1. HanrotF, SertD, DelinchantJ, PietruckR, BürglerT, BabichA, et al. In: Proceedings of 1st Spanish National Conference on Advances in Materials Recycling and Eco–Energy, Madrid, Spain, 2009:181–4.Search in Google Scholar

2. WakimotoK, NakamuraH, FujiiM, YamamotoY, NemotoK, TomiokaK. NKK Tech Rep1998;78:5965.Search in Google Scholar

3. SahajwallaV, HongL, Saha-ChaudhuryN. In: Proceedings of AISTech 2005, Charlotte, NC, USA, 2005:916.Search in Google Scholar

4. NomuraS, KatoK, NakagawaT, KomakiI. Fuel2003;82:177582.Search in Google Scholar

5. NomuraS, KatoK. Fuel2006;85:4756.Search in Google Scholar

6. BazarganA, McKayG. Chem Eng J2012;195–196:37791.10.1016/S0140-6736(11)60031-7Search in Google Scholar

7. MurakamiT, KasaiE. ISIJ Int2011;51:913.10.2355/isijinternational.51.9Search in Google Scholar

8. TurkdoganET. Physical chemistry of high temperature technology, New York, NY: Academic Press, 1980:528.Search in Google Scholar

9. Walker Jr P, VastolaF, HartP. In: H. Saltsburg, J. Smith and M. Rogers, editors. Fundamentals of gas-surface interactions. New York, NY: Academic Press, 1967:30717.Search in Google Scholar

10. AbrahamsonJ. Carbon1973;11:33762.Search in Google Scholar

11. LiK, KhannaR, ZhangJ, LiuZ, SahajwallaV, YangT, et al. Fuel2014;133:194215.Search in Google Scholar

12. HongL, SahajwallaV. Metall Mater Trans B2013;44:154145.10.1007/s11663-013-9920-ySearch in Google Scholar

13. KayacanI, DoganOM. Energy SourcesPart A2008;30:38591.Search in Google Scholar

14. EncinarJM, GonzalezJF. Fuel Process Tech2008;89:67886.10.1016/j.fuproc.2007.12.011Search in Google Scholar

15. MatsudaT, HasegawaM, IkemuraA, WakimotoK, IwaseM. ISIJ Int2008;48:118896.10.2355/isijinternational.48.1188Search in Google Scholar

16. ChoY-S, ShimM-J, KimS-W. Mater Chem Phys1998;52:947.Search in Google Scholar

Received: 2014-7-13
Accepted: 2014-12-29
Published Online: 2015-2-28
Published in Print: 2016-1-1

©2016 by De Gruyter

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 1.5.2024 from https://www.degruyter.com/document/doi/10.1515/htmp-2014-0122/html
Scroll to top button