Category Archives: Technologies for Converting Biomass to Useful Energy

Experimental setup and procedure

All co-firing experiments were conducted using a 30 kW (100,000 BTU/h, approximately 15 lb or 6.80 kg of coal/h) small-scale furnace capable of firing most types of ground fuels. A schematic of the furnace is shown in Figure 3.20. Propane and natural gas are used to heat the furnace to the operating temperature of 1100°C (2000°F). Type K (shielded, ungrounded) thermocouples are used to measure the temperature along the axial length of the furnace. A solid fuel hopper feeds coal and coal/biomass blends during experiments. Primary air (6m3/h, 15-20% of total air) is necessary to propel the finely ground solid fuel through the fuel line and to the furnace. Prior to ventilation, all exhaust gases pass through a water-cooling spray to significantly lower the temperature of the gases. A sump pump pumps this water out of the furnace. More details are provided in Frazitta et al. (1999), Arumugam et al. (2006c), Annamalai et al. (2005), Lawrence et al. (2009) and in Thien et al. (2012).

image115

Figure 3.19. Co-firing power plants in Europe (adopted from European Biomass Industry Association (EUBIA)).

image116

Figure 3.20. Schematic of boiler burner facility for co-firing (adapted from Lawrence, (2007).

The secondary air (75-80% of total air) heater was run for an hour before the experiment was started. The secondary air is swirled (Swirl number = 0.7) prior to entry into the combustion chamber. Once the secondary air reached a steady temperature, approximately 500 K (440.33°F), the propane torches were ignited. Natural gas and propane were used to preheat the furnace to operating temperatures. Once the furnace reached 1366 K (2000°F), the natural gas was turned off and the natural gas line was closed. The solid feeder line was opened and the solid feeder was turned on and set to the desired fuel flow rate. The primary and secondary air lines were set to the appropriate flow rates to obtain the desired equivalence ratio. The furnace was allowed to run for 30 minutes before the first readings were taken. The measurement was taken at the last sampling port just before the quenching water sprays and the wet flue gases were ducted to the atmosphere. After taking a measurement at this equivalence ratio, the secondary air could be adjusted to a different equivalence ratio. After taking measurements at all desired equivalence ratios, the furnace was turned off.

Fuel properties played a significant impact on the burnt fraction and the emissions created by combustion. The results from the co-firing experiments performed are discussed and their role in evaluating the combustion performance of the fuels is explained. The performance was evaluated by measuring combustion efficiency (burnt fraction) and the emissions levels of pollutants that include NOX and CO. In addition, overall fuel nitrogen conversion efficiency to NOX was also determined. The mercury emissions are presented elsewhere (Udayasarathy, 2007).

The co-firing involves a mix of finely pulverized biomass and coal. The size distribution was obtained using an ASTM sieve shaker. Sauter mean diameter (commonly abbreviated as SMD or d32) is commonly used for estimating the average size of solid fuel particles. The SMD is defined as the diameter of a sphere that has the same ratio of volume to surface area. It is represented as the following equation:

n

J2dl ■ n

SMD or d32 = -=————-

^ di ■ ni

i=1

where di is the diameter of particles and ni is the number of the particles of diameter di. According to the Rosin Rammler fit, the cumulative mass fraction CMF or drops (or particles) with a dimension lesser than dp is given as (Annamalai and Puri, 2007):

CMF = 1 — exp(-bdp)

where b: size constant, n: distribution constant and is a measure of spread of drop size. In terms

of the dp, charac size:

Подпись:CMF = 1 — exp

where dp, charac denotes the characteristic drop or particle size for which CMF = 1 — exp(-1) = 63.2% and

Подпись: b =1

dn

p, charac

Подпись: R = 1 — CMF = exp Подпись: dp dp,charac Подпись: n

The fraction R having size greater than dp is:

The plot of ln{R} vs. dp must be linear and the slope yields “n” and dp, charac is determined from the plot at R = 0.368. The values of “n”, “b” and SMD are presented in Table 3.9 for several fuels. Note that the coals had a larger SMD than those of DB fuels. The dirt (or mineral matter) that got collected with the DB fuels passed through all of the sieves and collected in the pan. This caused the DBs to have a smaller SMD.

Table 3.9. Size distribution parameters, adopted from Lawrence (2007).

Size distribution parameters

TXL WYO LA-PC-DB-SepS HA-PC-DB-SoilS

n 1.2991 1.4369 1.0934 1.2612

b 0.000934 0.00042 0.0024 0.0013

SMD (microns) 396 396 96.7 91.6

BIOMASS GASIFICATION IN CHINA

1.2.2 Introduction

Gasification is a technology commonly used nowadays for extracting energy from biomass. Over the past decade, there has been great progress in the development of gasification technology in China. Many kinds of biomass gasification processes have been developed, treating different materials for various purposes.

1.2.3 Gasification technology development

Dating back to the 19th century, gasification technology now attracts new interests in Europe, because of the end use flexibility of the syngas (Lan et al., 2011).

A charcoal gasifier had been developed in the 1940s and was tried to drive vehicles with the technology, which is the initial exploration in China in gasification of biomass. But the technology did not obtain further development for various reasons.

A fluidized bed reactor for industrial applications had been developed in China in the 1950s, but there were some imperfections in the technology, and the application was suspended. In the 1960s, Chinese researchers began to study the biomass gasification power generation and gained some experiences, and the preliminary prototype had been developed and gained some experiences, but these researches were stopped because of the economic conditions and the small profits.

A fixed bed gasifier (updraft and downdraft) circulating fluidized bed gasifer was developed in China in the 1980s. The product gases were used for power generation, supplying heat and cooking.

An 1 MW BGPG system with a circulating fluidized bed (CFB) gasifier had been developed, and constructed in a rice mill in the Fujian province of China (Wu et al., 2002; Wang et al., 2005; Lu et al., 2004b; Wu et al., 2003). A neural network was focused on for the simulation of gasification process. An artificial neural network model was developed to simulate the gasification processes in order to obtain the gasification profiles (Guo et al., 2001; Tang et al., 2003; Wang et al., 2002).

Gasification and polygeneration technology in the fluidized bed were concentrated on by Tie et al. (2003; 2005). They studied four kinds of biomass (bagasse, pine sawdust, peanut shell, rice husk) in a fluidized bed reactor and found that the temperature was a key parameter due to biomass pyrolysis in a fluidized bed reactor.

Lu etal. (2005; 2007) developed hydrogen production technology by using supercritical water. He and other researchers (Guo et al., 2006; Yan et al., 2006) mixed several kinds of biomass in sodium carboxymethylcellulose, which was gasified successfully at 650°C, 25 MPa in a tubular flow reactor with formation of hydrogen, carbon dioxide, carbon monoxide, methane and a small amount of ethane and ethylene.

Furthermore, the domestic garbage gasification was studied by Yuan et al. (2002). Chen et al. (2003a, b; 2005; 2006; 2008) made progress in cogasification of biomass and glycerin. They studied four kinds of biomass in a two-stage reactor to produce hydrogen-rich gas, and investigated the effect of a catalytic bed on the pyrolysis behavior.

Chemical looping reactions

In many chemical looping processes calcium sorbents are used to separate CO2 from the other gases. These reactions generally involve absorption/calcination processes, based on the following

Подпись: Hot air

image291 Подпись: Oxidizer

image293CO? and SO}

Figure 5.28. The Alstom process (Rizeq et al., 2002; Andrus et al., 2006).

chemical reactions:

CaO + CO2 ^ CaCO3

+heat (5.36)

CaCO3 —CaO + CO2

These reactions can be shifted to the left or to the right by changing the temperature and pressure.

The combustion of carbonaceous fuels can be assumed as follows:

(lx + y/2 — z) MO + CxHyOz ^ (lx + y/2 — z)M + xCO2 + y/2H2O

M + Air ^ MO + N2 + unreacted O2 (5.37)

where the looping medium is a metal oxide. The first reaction only produces CO2 and steam; therefore carbon dioxide may be easily separated from steam by condensing the latter. The second reactor oxidizes the metal again and the flue gases are nitrogen and oxygen. The second reaction provides heat to the first reactor. The overall combustion process is thus divided into two sub processes which separate inherently the flue gas components.

If the fuel consists of carbon only, then the chemical looping reactions can be as follows:

Подпись:2MO + C ^ 2M + CO2 M + H2O ^ MO + H2

The looping medium is the same as in the previous reactions, but the flue gas of the second reactors is now pure hydrogen. The process allows separating CO2 for capture and producing hydrogen as a byproduct.

The second reactor produces hydrogen instead of heat and steam is the oxidant of the overall combustion process. The carbon capture is greatly simplified and requires much less energy than in any other process producing hydrogen.

Hydrogen can also be produced including a calcination process in a chemical looping combustion:

Подпись:CO (g) + H2O (g) ^ CO2 (g) + H2 (g) CO2 (g) + CaO (s) ^ CaCO3 (s) CaCOs (s) ^ CO2 (g) + CaO (s)

One positive effect of capturing CO2 in the process by means of CaO, is to enhance the hydrogen production from the first reaction because the chemical equilibrium favors its formation. The calciner in the second reactor allows an efficient separation of CO2 from the flue gas which can be ready for capture.

High pressure and turbulent syngas flames

There has been relatively little work concerning high pressure syngas flames. McLean et al. (1994) and Vagelopoulos and Egolfopoulos (1998) reported premixed flame speeds at pressures from atmospheric to a few atmospheres. Burke et al. (2007) examined the effect of CO2 on burning velocity of spherically expanding flames at p = 1.0 and 10 atm using a 25%H2-75%CO mixture with 12.5%O2-87.5%He oxidizer. Sun et al. (2005) reported laminar flame speeds for

image028

Diffusion Ф=5 Ф=2 Ф=1.6 Ф=1.0

Figure 2.15. Images of laminar partially premixed 45%H2/35%CO/20%CO2-air flames at different levels of partial premixing and Reynolds number of 1400 (Ouimette and Seers, 2009).

CO/H2/air and CO/H2/O2/He mixtures for pressures up to 40 atmospheres using the constant- pressure spherical flame technique. A kinetic model was also developed using the latest available thermo-transport and kinetic data (Park et al., 2004; Ouimette, 2009). The mechanism was validated against the measured flame speeds, non-premixed counter flow ignition temperatures, concentration profiles in a flow reactor, and ignition data from shock tube experiments. Figure 2.16 from their study shows the measured and predicted laminar flame speeds plotted versus Ф for CO/H2/O2/He mixtures at different CO/H2 ratios, and pressures of 5-40 atm.

Predictions are based on their kinetic model and that reported by Davis et al. (2005). As expected, the flame speed increases with increasing H2 content, and decreases with increasing pressure. Overall, there is good agreement between the predictions and measurements, although both models exhibit discrepancies, whichmay be attributed to uncertainties in kinetic and transport data. Thus, further studies are warranted for high-pressure syngas flames over a range of com­bustion regimes, including non-premixed and partially premixed combustion, and using different burners.

Studies on turbulent syngas flames have focused on the determination of turbulent flame speeds (ST) (Chase et al., 1951; Kee et al., 1995; Daniele, 2011). While ST can be defined in multiple ways, it is often based on a global consumption speed (Venkateswaran et al., 2011) and is presented in terms of the normalized flame speed (ST/SL) as a function of turbulence intensity, fuel composition and other parameters. Daniele et al. (2011) considered the reaction zones regime and examined the effects of pressure and syngas composition on the turbulent flame speed. Correlations were developed for ST / SL as a function of normalized parameters representing the effects of turbulence intensity, integral length scale, pressure, and temperature.

The increase of ST/SL with increasing pressure and H2 content was attributed to the thermo­diffusive and hydrodynamic instabilities. Venkateswaran et al. (2011) reported measurements of global turbulent flame speeds using a Bunsen burner, and examined the effects of Ф, syngas composition, mean flow velocity, and turbulence intensity. Consistent with other studies, the flame speed was found to exhibit sensitivity to fuel composition over a wide range of turbulence intensity, increasing significantly with the increase in H2 content. The data were further analyzed to develop flame speed correlations, indicating the effects of thermo-diffusive instabilities through negative Markstein lengths.

image029

Figure 2.16. Measured and predicted laminar flame speeds versus Ф for different СО/Н2/НЄ/О2 mixtures at 5, 10, 20, and 40 atm. Predictions are based on the kinetic models of Sun et al. (2005) (solid line) and Davis et al. (2005) (dashed line).

Experimental procedure

A normal experiment started with preheating the grate and the combustion chamber using a propane torch placed under the grate. When the temperature in the combustion chamber (2 cm above the grate) reached 800°C (after ~2 hours), the torch was turned off and biomass was added to the gasifier. The addition continued until the bed height attained 17cm; afterwards, the fuel port was closed and the flows of steam and air were adjusted to the desired experimental conditions. As the biomass was pyrolyzed and the char was burned the bed height started decreasing and the ash accumulated. Thus, biomass was added every 10 minutes and in batches as required. In the earlier batch experiments reported by Priyadarsan et al. (2004), there was no ash disposal system; as such the temperature peak moved towards the bed surface due to ash accumulation at the bottom. In the current experiments, the ash was disposed of continuously and a quasi-steady state was assured by maintaining the peak temperature at the same location in the ash disposal system. When the peak temperature achieved a steady state (~1.0 hours) the gas sampling unit was turned on and the gas analysis was performed continuously during 20 minutes by the mass spectrometer (MS).

The flowrate of dairy biomass was maintained constant at 1 kg/h and the flows of air (0.56-2.26 SATP m3/h (standard ambient temperature and pressure meter cube per hour)) at 15°C and steam (0.19-0.43 kg/h) at 100°C were changed in order to obtain the desired experimental conditions: ER = 1.59, 2.12, 3.18, 4.24, and 6.36 and S:F = 0.35, 0.56, 0.68, and 0.80. An air drier was used to dry the air before it was supplied to the gasifier. The gasifier was operated at 98 Pa vacuum pressure during all the experimentation. Temperatures along the gasifier were monitored at every 60 seconds by type K thermocouples located at 0.02, 0.04, 0.07, 0.13, 0.20, 0.24, and 0.28 m above of the grate. Samples were taken at the top of the gasifier at the rate of 0.14 SATP m3/h and conditioned by the sampling unit in order to remove tar and particulate material. The mole fractions of CO2, CO, CH4, C2H6, O2, H2, and N2 were measured every ten seconds by the MS. The same procedure used for the air gasification was again employed for enriched air gasification with little changes.

Biomass composition and analysis

To evaluate which conversion process is more suitable for different biomasses, a preliminary characterization is necessary to analyze their chemical, physical and energetic properties.

Biomass is composed of water, ashes and dry matter without ashes and only the latter component is interesting for energy conversion yielding a calorific value. Ashes and water decrease the commercial value of biomass because:

• they decrease the bulk energy content of biomass;

• moisture absorbs energy for evaporation;

• ashes have to be disposed of;

• light ashes are transported by flue gases and contribute to PM (particulate matter) emissions;

• low melting point ashes foul heat exchangers.

Given the presence of these three different main components, the measurable quantities contained in a biomass can be expressed (Fig. 5.2): [6]

Table 5.1. Selected biomass characteristics. VM: volatile matter, HHV: higher heating value, LHV: lower heating value (Mancosu, 2011)

Bulk

Biomass

Moisture [%] wb

VM

[%]

Ashes

[%]

C

[%]

H

[%]

N

[%]

HHV

[MJ/kgdb]

LHV

[MJ/kgdb]

density

[kg/m3]

Oak wood

6.2

86.0

0.9

49.7

6.5

0.2

20.4

18.9

750

Pine wood

9.5

89.3

0.7

51.3

6.1

0.2

19.2

440-560

Pine bark

1.8

46.9

5.3

Pellet

10.0

85.6

0.8

49.8

6.4

0.3

18.5

17.4

650

Sorghum

2.1

43.9

6.2

0.2

16.8

220-260

Salix wood

7.9

85.7

1.9

49.1

6.2

0.3

18.8

300-400

Poplar wood

8.6

80.3

1.3

49.7

6.5

0.2

19.6

19.3

420

Fire wood

7.7

77.0

5.8

48.6

6.5

0.2

18.9

700-800

Birch wood

7.4

80.9

2.6

48.3

8.3

0.1

19.3

600

Vine pruning

45.0

86.0

2.3

46.5

6.4

0.4

18.6

17.1

790-900

Olive tree pruning

40.0

86.0

3.9

49.3

5.5

0.6

18.5

17.4

800-900

Sawdust

11.6

81.5

0.8

49.5

6.8

0.4

19.7

100

Bamboo

8.5

76.5

0.8

50.6

5.3

0.2

19.3

200-250

Wood chips

9.3

88.0

1.0

50.0

5.8

0.3

19.3

150

Giant reed

40.0

8.5

45.5

5.7

0.2

18.0

17.5

180-200

Black locust

30.0

85.7

3.6

50.7

5.7

0.5

19.7

18.5

625

Straw

8.7

72.3

14.9

43.0

6.3

0.8

16.0

14.9

100-180

Wheat

6.4

75.0

8.0

43.0

10.85

0.3

16.0

Rice husk

69.3

19.0

36.7

5.0

0.9

14.5

13.9

75

Sugarcane

85.2

2.2

52.5

6.8

0.5

18.9

130-150

Rapeseed

6.1

77.6

3.8

42.4

7.1

0.2

16.6

Stone fruit resid.

6.9

85.6

0.5

51.6

6.0

0.5

21.6

Almond shell

8.7

81.7

2.8

52.4

6.7

0.5

19.0

17.7

Hazelnut shell

9.3

71.0

7.9

42.8

5.15

0.6

15.7

Walnut shell

6.7

76.1

3.6

51.5

7.3

0.7

Tomato

7.0

86.1

3.8

52.3

7.6

3.4

Olive husk

8.3

78.4

6.4

49.6

5.5

1.4

20.9

19.1

Bagasse

77.7

2.1

51.5

6.0

1.0

18.2

Ultimate analysis is defined as “the determination of the elemental composition of the organic portion of carbonaceous materials, as well as the total ash and moisture” (Miller and Tillman, 2008; ASTM D 5373-02; Milne et al., 1990).

Fluidized bed gasifiers

The basis for the fluidized bed reactor configurations is the principle of fluidization. By forcing a gas stream (fluidization medium) through a reactor, the fuel together with the inert bed material will behave like a fluid, if the flow velocity is high enough. Air, steam or steam/oxygen mixtures are examples of commonly used fluidization media. Silica sand is the most extensively used bed material, but other bulk solids, especially those exhibiting a catalytic activity, such as olivine sand and dolomite, are also employed.

Fluidized beds provide many features not available in the fixed-bed types, including high rates of heat and mass transfer and good mixing of the solid-phase, which means that reaction rates are high and the temperature is more or less constant in the bed.

Depending on the velocity of the fluidization medium, the fluidized bed gasifiers may be divided into two categories, bubbling fluidized bed (BFB) gasifiers and circulating fluidized bed (CFB) gasifiers.

Empirical formula for heat values

3.2.2.1 The higher heating value per unit mass of fuel

The gross or higher heating values HHV for coals can also be empirically obtained by using the Dulong equation (Annamalai and Puri, 2007), namely:

HHV[kJ/kg] = 33800 YC + 144153 YH — 18019 YO + 9412 YS (3.2)

where YC, YH, YN, YO and YS are mass fractions of C, H, N, O and S.

Another relation due to Mott and Spooner is (Mason and Gandhi, 1980):

if O < 15%

HHV [kJ/dry kg] = 103.5C%+ 1418.3 x H% + 94.2S% — 145.1 x O (organic)% (3.3) if O > 15%

HHV [kJ/dry kg] = 103.5 x C% + 1418.3 x H% + 94.2 x S%

— {153.2 — 72 x O%/(100 — A%)} x O% (3.4)

Here A = ash content.

Channiwala (1992) considered over 200 species of biomass and fitted the following equation to the data:

HHV [kJ/dry kg] = 34910 YC + 117830 YC — 10340 YO — 21110 YA + 10050 YS — 1510 YN

(3.5)

The experimental data have an error of about 1.5%.

Boie empirical equation for HHV of any fuel CcHhNnOoSs (Annamalai and Puri, 2007):

HHV[kJ/kmole] = 422272 x C + 117387 x H — 155371 x O + 100480 xN + 335508 x S (3.6)

where C, H, O, N and S are the number of carbon, hydrogen, oxygen, nitrogen and sulfur atoms respectively in the fuel. The same equation can be used to determine the stoichiometric oxygen in kg per empirical kg of fuel:

Vo2 = 32 {C + H/4 — (1/2)O + S} = 32C{1 + (H/C)/4 — (1/2)(O/C) + (S/C)} (3.7)

HHV [kJ/kg] = C{422272 + 117387 x (H/C) — 155371 x (O/C)

+ 100480(N/C) + 335508 x (S/C)} (3.8)

Based on the Boie equation, the enthalpy of formation can be derived as:

h0FJ = 28752 x{C — 0.888 x H — 6.168 x O + 6.199N + 1.337 S} [kJ/kmole] (3.9)

Table 3.3. Fuel Properties (adopted from Sweeten et al., 2006 and TAMU, 2006).

Fuel

Type

Source

Ash

Dry loss

FC

VM

C

Coal

ar

T1: Fuel Properties

5.3300

15.1200

42.3800

37.1700

60.3000

Litter biomass

ar

T1: Fuel Properties

26.8100

11.6200

10.9100

50.6500

28.4400

Sewage sludges in Thailand (C1)

dry

Predicting the heating values of sewage sludges in Thailand

38.4000

6.1000

8.6000

53.0000

31.1000

Sewage sludges in Thailand (C2)

dry

Predicting the heating values of sewage sludges in Thailand

42.0000

5.1000

6.7000

51.2000

27.5000

Sewage sludges in Thailand (C3)

dry

Predicting the heating values of sewage sludges in Thailand

43.0000

5.4000

7.0000

50.0000

26.4000

Sewage sludges in Thailand (C4)

dry

Predicting the heating values of sewage sludges in Thailand

48.4000

6.4000

4.0000

47.6000

23.9000

Sewage sludges in Thailand (C5)

dry

Predicting the heating values of sewage sludges in Thailand

51.8000

3.7000

6.0000

42.2000

20.9000

Sewage sludges in Thailand (C6)

dry

Predicting the heating values of sewage sludges in Thailand

61.8000

4.1000

3.7000

34.5000

18.0000

Sewage sludges in Thailand (C7)

dry

Predicting the heating values of sewage sludges in Thailand

56.0000

3.4000

5.0000

39.0000

19.5000

Sewage sludges in Thailand (C8)

dry

Predicting the heating values of sewage sludges in Thailand

63.5000

3.9000

3.2000

33.3000

14.5000

Sewage sludges in Thailand (C9)

dry

Predicting the heating values of sewage sludges in Thailand

64.0000

3.7000

3.1000

32.9000

15.3000

Sewage sludges in Thailand (C10)

dry

Predicting the heating values of sewage sludges in Thailand

67.6000

3.2000

1.8000

30.6000

12.7000

Sewage sludges in Thailand (C11)

dry

Predicting the heating values of sewage sludges in Thailand

72.9000

4.4000

2.2000

24.8000

10.6000

Sewage sludges in Thailand (H1)

dry

Predicting the heating values of sewage sludges in Thailand

39.4000

6.6000

5.1000

55.5000

26.7000

Sewage sludges in Thailand (H2)

dry

Predicting the heating values of sewage sludges in Thailand

40.6000

5.6000

6.8000

52.6000

29.6000

Sewage sludges in Thailand (H3)

dry

Predicting the heating values of sewage sludges in Thailand

45.9000

4.6000

6.5000

47.7000

25.5000

Sewage sludges in Thailand (H4)

dry

Predicting the heating values of sewage sludges in Thailand

45.7000

6.9000

3.9000

50.4000

25.0000

Sewage sludges in Thailand (H5)

dry

Predicting the heating values of sewage sludges in Thailand

60.2000

4.6000

3.2000

36.6000

19.0000

Sewage sludges in Thailand (I1)

dry

Predicting the heating values of sewage sludges in Thailand

42.3000

5.2000

3.2000

54.5000

25.1000

Sewage sludges in Thailand (I2)

dry

Predicting the heating values of sewage sludges in Thailand

51.6000

5.0000

2.8000

45.6000

22.6000

ar: as received

HHV-DAF

H

N

O

S

HHV,

kJ/kg

Boie,

kJ/kg

HHV-DAF,

kJ/kg

Chemical formula

3.6200

0.9600

14.5000

0.2300

23710

30025

29805

CH0.7139N0.0136O0.1805S0.0014

3.7100

3.0350

22.7900

0.6600

12060

19564

19587

ch1.5512n0.0915°0.6015s0.0087

4.2000

3.3000

24.3000

1.1000

13900

21824

22565

CH1.6059N0.0910O0.5865S0.0132

4.1000

4.0000

23.3000

1.1000

13200

21099

22759

ch1.7729n0.1247o0.6360s0.0150

4.1000

4.3000

23.7000

0.9000

12600

20673

22105

ch1.8467n0.1396o0.6739s0.0128

3.9000

3.8000

21.8000

1.3000

11000

21111

21318

CH1.9404N0.1363O0.6847S0.0204

3.4000

3.3000

21.7000

0.9000

10100

19077

20954

ch1.9344n0.1354o0.7794s0.0161

2.9000

2.3000

16.7000

0.8000

9400

21140

24607

ch1.9158n0.1095o0.6964s0.0166

3.2000

3.1000

19.4000

0.8000

8700

19778

19773

CH1.9514N0.1363O0.7468S0.0154

2.6000

2.6000

18.1000

1.2000

6900

17539

18904

CH2.1322N0.1537O0.9370S0.0310

2.5000

2.3000

17.7000

0.5000

6500

18108

18056

ch1.9430n0.1289o0.8684s0.0122

2.0000

1.8000

17.5000

0.6000

5700

15509

17593

CH1.9430N0.1289O0.8684S0.0122

2.0000

1.6000

15.7000

0.4000

4300

16491

15867

CH2.2436N0.1294O1.1118 S0.0141

4.0000

4.3000

27.5000

0.7000

13300

18697

21947

CH1.7814N0.1381O0.7731 s0.0098

4.6000

5.0000

21.5000

1.0000

12800

23212

21549

ch1.8479n0.1448o0.5452s0.0127

3.9000

4.2000

21.7000

1.0000

12400

21145

22921

CH1.8186N0.1412O0.6388S0.0147

3.8000

3.7000

24.3000

0.8000

11100

19941

20442

ch1.8074n0.1269o0.7296s0.0120

3.0000

2.7000

16.8000

1.2000

8200

21606

20603

ch1.8775n0.1218o0.6637s0.0237

4.0000

3.8000

26.1000

0.9000

10900

18912

18891

CH1.8950N0.1298O0.7805S0.0134

3.2000

2.9000

20.3000

2.0000

9900

20259

20455

ch1.6837n0.1100o0.6742s0.0332

Table 3.3. Continued.

Fuel

Type

Source

Ash

Dry loss

FC

VM

C

Sewage sludges in Thailand (I3)

dry

Predicting the heating values of sewage sludges in Thailand

58.8000

4.7000

3.0000

38.2000

18.3000

Misc. manure

ar

n/a

36.3825

50.5000

2.3760

10.7910

9.7020

Sheep manure

ar

n/a

10.9098

47.8000

7.3080

34.0344

21.1932

Mortality Biomass source:

ar

Properties ofthe fuels (MB)

34.2000

0.9600

10.4700

54.3700

38.4500

Brent

Auvermann

(before

treatment)

Cofired coal

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

14.7000

5.0000

61.1400

24.1600

72.7500

Pine shavings

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

0.1000

45.0000

15.2000

84.7000

49.1000

Reed Canary Grass

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

4.1000

65.2000

19.8000

76.1000

45.8000

Sheep manure

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

20.9000

47.8000

14.0000

65.2000

40.6000

Dairy free-stall

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

62.3000

70.3000

7.1000

30.6000

22.1000

Misc. manure

dry

T3: Proximate, ultimate and ash analyses of coal, pine shavings, and animal waste

73.5000

50.5000

4.8000

21.8000

19.6000

DB soil surface

ar

n/a

59.9100

12.2100

3.9200

23.9900

18.0200

DB seperated

ar

n/a

14.9300

25.2600

13.0000

46.8800

35.2000

solids

Texas lignite

ar

n/a

11.5000

38.3000

25.4000

24.8000

37.2000

Wyoming

ar

n/a

5.6000

32.9000

33.0000

28.5000

46.5000

sub-bituminous

Thus an approximate method based on the Boie heat value exists to compute hf of any empirical fuel. If only mass fractions of C, H, N, O and S are known as YC, YH, YN, YO and YS, then the higher heating value of the fuel becomes:

HHVF [kJ/kgfuel] = 35160 YC + 116225 YH — 11090 YO + 6280 YN + 10465 YS (3.10)

One can deduce the lower or net heat value (LHV) when hydrogen in water is excluded giving: LHVf [kJ/kgfuel] = 35160 YC + 94438 YH — 11090 YO + 6280 YN + 10465 YS (3.11)

HHV-DAF

H

N

O

S

HHV,

kJ/kg

Boie,

kJ/kg

HHV-DAF,

kJ/kg

Chemical formula

3.4000

1.8000

18.7000

1.8000

9000

20907

21845

CH2.2093 N0.0843 O0.7670 S0.0368

1.2375

0.4950

1.6335

0.0495

3585

35730

27330

CH1.5167N0.0437°0.1264 S0.0019

2.6622

1.0962

16.0254

0.3132

8372

21455

20275

CH1.4937N0.0443 O0.5676 S0.0055

3.9700

0.2900

22.7100

0.0500

12774

24118

19701

CH1.2278N0.0065O0.4433 S0.0005

3.9100

1.5000

4.8700

2.2700

30512

35070

35770

ch0.6391n0.0177o0.0502s0.0117

6.4000

0.2000

44.0000

0.2000

19475

19876

19494

CH1.5500N0.0035 O0.6727 S0.0015

45.8000

1.0000

42.9000

0.1000

16838

19300

17558

CH1.5837N0.0187O0.7031 s0.0008

5.1000

2.1000

30.7000

0.6000

16037

21455

20274

ch1.4937n0.0443 O0.5676 S0.0055

2.9000

1.1000

11.5000

0.1000

8836

26380

23438

CH1.5604N0.0427O0.3906 S0.0017

2.5000

1.0000

3.3000

0.1000

7243

35730

27332

CH1.5167N0.0437O0.1264 S0.0019

1.4500

1.1500

7.0700

0.1900

4303

26260

15434

CH0.9568N0.0547O0.2945 s0.0039

3.1200

1.9300

19.1500

0.4300

12817

23455

21430

CH1.0540N0.0470O0.4084 S0.0046

2.1000

0.7000

9.6000

0.6000

14290

29009

28466

CH0.6713N0.0161 O0.1937 S0.0060

2.7000

0.7000

11.3000

0.3000

18194

29772

29584

CH0.6905 N0.0129O0.1824 S0.0024

Correlation for adiabatic flame temperature with ash and moisture content is shown and plotted in Figure 3.8.

Figure 3.9 shows the higher heat or gross heat value of C-H-O fuel in kJ per kg of fuel.

3.2.2.2

image042 Подпись: HHV VO2 Подпись: (3.12)

The higher heat value per unit stoichiometric oxygen The heat value per unit stoichiometric oxygen (vO2) defined as:

image045

Figure 3.6. Synergistic NOX reduction from co-firing biomass (adopted from Tillman, 2000).

image046

Figure 3.7. Higher heating valuesHHV for cattle ration, raw FB, partially composted FB, finished com­posted FB, coal, and respective FB + 5% crop residue blends (adopted from Sweeten et al., 2003).

It is well known that the HHVO2 is almost constant for most fuels. For Boie equation, the HHVO2 is given as:

HHVO2 [kJ/kg of O2] = {422272 + 117387 x (H/C) — 155371 x (O/C) + 100480(N/C)

+ 335508 x (S/C)}/(32{1 + (H/C)/4 — (1/2)(O/C) + (S/C)})

(3.13)

image047

Figure 3.8. Correlation of adiabatic flame temperature with moisture and ash contents; Tadiabatic [K] = 2285 — 1.8864 x H2O + 5.0571 x Ash — 0.3089 x H2O x Ash — 0.1802 x H2O2 — 0.1076 x Ash2, H2O and Ash in fractions; multiply T adiabatic in K by 1.8 to obtain T (Annamalai et al., 2007b; Sami et al., 2001).

image048

Figure 3.9. Variation ofHHV with H/C and O/C in C-H-O fuels.

Ignoring S and N, trace elements in fuel, Figure 3.10 plots HHVo2, in kJ per kg of oxygen as HHV/vO2 constant. It is apparent that the HHV per unit mass of O2 burned is approximately the same of about 14250 kJ/kg of oxygen (18.6kJ/SATP liter of oxygen, where SATP means at standard atmospheric temperature and pressure ) or 3280kJ/kg stoich air (3.9kJ/SATP liter of air) for most fuels. For methane, the literature states that HV per unit O2 is 13550 kJ per kg of O2 (17.7kJ/SATP liter of O2) while Boie based equation yields 13934 kJ/kg of O2. For n-octane,

image049

image050the value is 13640 kJ per kg of O2 or 17.82 kJ/liter of O2 (SATP) for CH4 while Boie yields 13730 kJ/kg O2 for Octane.

Figure 3.11 plots the respiratory quotient (RQ), a term used in biological literature (Annamalai and Silva, 2011) and defined as CO2 per kmole of stoichiometric oxygen, an indication of global

warming potential) for various biomass fuels. Typically RQ is about 1 (which is same as that of glucose, C6Hi2O6) for biomass fuels.

3.2.2.3 Heat value of volatile matter

In Figure 3.11 we see how H/C relates in fat, protein, biomass and coal. If the heat of pyrolysis is neglected, the heat of combustion of the coal can be represented as a combination of the contri­bution from the volatile matter (HVVVM) and the contribution from the fixed carbon (HVCFC) in relation to their mass percentages:

Подпись: (3.14)HV Coal = HVv VM + HVc FC

image052 Подпись: (3.15)

If FC = 1 — VM as in the case of dry ash-free (DAF) coals, one can correlate the heating values of volatiles HVv to VM (Annamalai and Puri, 2007). The Volatile Matter Higher heating value (HHVVM) was calculated using:

where HHV is the as received heating value, FC% is the amount of fixed carbon present in the fuel, HHVFC is the higher heating value of the fixed carbon (enthalpy of formation/molecular weight), and VM% is the amount of volatile matter present in the fuel.

Advantages and disadvantages

Co-combustion in large-scale power plants can lead to an overall saving of fuels in comparison to independent fossil — and biomass-fired plants. Also, it can increase the fuel flexibility and reduce investment cost. Comparing with coal, biomass is a renewable energy source, which is considered as a CO2-neutral fuel with lower emissions of SO2, NOx, heavy metals. NOx emissions could be reduced by biomass, which has a low nitrogen and high volatile content. (EUBIA, 2007; Zhang et al., 2010; Fu et al., 2009; Daniele et al., 2007). The co-combustion of coal and biomass has many advantages which can be described as follows (EUBIA, 2007; VGB, 2008):

1. Reducing greenhouse gases emission — biomass is considered as a ‘carbon neutral’ fuel in that the CO2 emitted during biomass combustion is equal to that absorbed during the biomass growing. When biomass displaces a fossil fuel, a net reduction in greenhouse gas emissions is achieved.

2. Reducing local air pollutant emissions — burning biomass instead of fossil fuel results in lower emissions of SO2 and NOx.

3. Increasing electrical efficiency — the electrical efficiency of co-combustion power plant is higher than the traditional biomass plant, which has a small scale.

4. Ensuring security of supply — there exists a wide range of usable biomass fuels. Varying qualities and quantities of fuels can be partially compensated by adjusting the co-combustion rate.

5. Reducing cost — co-combustion presents the opportunity to use the existing fossil-fuel fired power plant infrastructure, which can be modified for co-combustion relatively easily. An optimum thermal biomass blending ratio of biomass co-combustion is 10% (on an energy basis) (Munir et al., 2010). Addition of biomass to a coal-fired boiler does not impact or at worst only slightly decreases the overall generation efficiency of a coal-fired power plant. Compared with other renewable options, biomass co-firing represents the most cost-effective means of renewable power generation in many cases (Belosevic et al., 2010; Baxter et al., 2005; Hein etal., 1998).

Meanwhile, co-combustion of coal and biomass has some disadvantages shown as follows (VGB, 2008):

1. Preprocess — A fuel handling system is designed for a particular water content, size distribution, dust etc. With co-combustion of biomass it is necessary to adapt the existing or even build a new combustion system for that fuel.

2. Corrosion — Higher corrosion risk due to increased HCl formation in case of substitution of fuels with higher chlorine content (sewage sludge, some cereals). Many biomass fuels contain large amounts of alkalines, especially potassium, which may aggravate the fouling problems (Baxter et al., 1993; Bakker et al., 1997; Robinsin et al., 2001a, b; Dunaway et al., 2003; Lokare etal., 2003).

3. A SCR DeNOx catalyst can be blocked by ash particles or deactivated by potassium, chlorine, and in case of sewage sludge also poisoned by some heavy metals and metalloids (As, Zn).

4. Operating costs are typically higher for biomass than for coal. The most sensitive factor is the fuel cost. Even if the fuel is nominally free at the point of its generation (as many residues are), its transportation, preparation and on-site handling typically increase its effective cost per unit energy such that it rivals and sometimes exceeds that of coal.

For the utilization of the ash in the cement and concrete industry, the concentrations of alkali metals, P2O5, SO3, Cl and unburned carbon in the ash are the critical parameters. It was found that the ashing temperature should be selected according to the biomasses proportion, when the biomass fraction is raised, the ash fusing temperature of blends decreases generally, and biomass with high P and K content proportion should not exceed 10% in co-firing (Dong et al., 2010).

Volatiles oxidation (flaming combustion)

This process originates the visible flame following the reactions of flaming combustion. Oxidation happens generating a diffusive flame due to slow combustion because volatiles substances exiting form the particle and oxidizing agent are not pre-mixed.

As reported by (Miller and Tillman, 2008) during volatile oxidation different free-radical reactions happen, such as: [7]

image232

Figure 5.9. Thermogravimetric curve for biomass (Biagini and Tognotti, 2006).

Table 5.4. Char combustion rates for different biomasses (Biagini and Tognotti, 2006).

Fuel

Peak

E [MJ/kmol]

A [min

—1]

Wood pellet

d1

69

6.70 x

105

Wood pellet

d2

346

6.96 x

1029

Wood pellet

ch

107

1.32 x

07

Pine wood

d1

75

1.62 x

106

Pine wood

d2

344

2.90 x

1029

Pine wood

ch

129

5.05 x

08

Cacao residue

d1

146

7.33 x

015

Cacao residue

d2

51

1.80 x

104

Cacao residue

ch

120

9.20 x

107

A simple chemical equation for levoglucosan combustion is the following (Sullivan and Ball, 2012):

C6H10O5 + 6O2 ^ 6CO2 + 5H2O (5.18)

Besides the stoichiometric equation it has to be considered that several intermediate compounds are created during flaming combustion of the volatiles. Woodley (1971) has identified about 40 compounds, produced by the thermal degradation of levoglucosan.

The oxidation reactions are exothermic and very fast. The activation energy for the oxidation of levoglucosan is about 190kJ/mol, the frequency factor is about 2.55 + 1013 s—1. The reaction enthalpy for complete combustion is —14kJ/g (Parker and LeVan, 1989).

Some other gas phase homogeneous reactions are reported in the following Table 5.5.