Sei sulla pagina 1di 5

16th European Conference & Exhibition on Biomass for Energy, Biomass Resources, 2-6 June 2008, Valencia, Spain

DRYING COSTS OF WOODY BIOMASS IN A SEMI-INDUSTRIAL EXPERIMENTAL ROTARY DRYER J. Meza*, A. Gil, C. Corts, A. Gonzlez CIRCE (Centre of Research for Energy Resources and Consumption) C/Maria de Luna, 3, 50018 Zaragoza, Spain *corresponding author: jmeza@unizar.es; tel: +34 976762562; fax: +34 976732078

ABSTRACT: Co-firing biomass in conventional coal-fired power plants seems to be a suitable and economical alternative to increase renewable energy share and accomplish CO2 reduction targets. However, the process may face up to some technological problems, such as excessive pre-treatment needed to burn biomass and therefore, a previous economic evaluation will be essential to determinate the optimal pre-treatment level given to justify the investment. Even more, recent application of biofuels as densified products (i.e., pellets and briquettes) and its raw material should accomplish final requirements of moisture and size. Actually, in industry, different types of dryers (flash, cascade, steam and rotary) are available for this purpose. However, it is known that rotary dryers have low cost of maintenance and consume 15% and 30% less specific energy that the flash and cascade types, respectively. Within this context, this paper evaluates the forest residue drying energy requirements in a semi-industrial experimental cocurrent flow rotary dryer. The apparatus is equipped with standard instruments for measuring air flow, temperature and humidity. These results will be the starting point for scaling-up of specific energy according to different biomass types and size. Thermal and power consumptions were split and considered for analysis. Keywords: biomass drying, forest residues, rotary kiln, wood chips.

INTRODUCTION

Recent policies across the OECD countries will call for an estimated additional 132000 GWh of biomassbased electricity by 2010, as essential target to produce about 7.7 million TJ of bioenergy [1]. These figures represent a 30% increase in bioenergy requirements for industrial and residential purposes. The growing use of bioenergy is too likely to increase forestry feedstock/timber/chip costs, which would have a future consequence on the energy sector and the traditional forestry sector. This is a challenge that the governments and the world should assume. In Spain, according to recent data, biomass amounted to 3% of primary energy consumption, or 4.5 Mtoe over a total of 145 Mtoe, contributing to renewable activity close to 7% [2]. However, only 22% of this was used to produce electricity. The rest of the consumption was thermal biomass. Even this way, according to the Plan to Promote Renewable Energy Sources in Spain 20052010, electrical biomass part is not even enough. For example, 4070 ktoe/year of thermal biomass were projected by 2010 and nowadays the amount is already 3536, whereas only 20% of the objective of produced electric biomass (409 of 2039 MWe) has been accomplished. The reasons are diverse, however, although the electric production from renewable sources is regulated since 1980, fortunately new legislation (May 2007) has taken place. The changes constitute, e.g., a possible incentive to co-firing in ordinary regulation power plants, or economic incentive in proportion to the primary renewable energy used in CHP stations (combined heat and power). In this order of ideas, the Spanish future perspectives with regard to use of biomass are increasing. Nowadays, densified biofuels (i.e., wood pellets and briquettes) are an ideal transport method from the forest to distant processing facilities. Wood pellets are generally produced from wood waste such as sawdust and thus should be viewed as part of a complete forest products manufacturing sector. The raw material is dried, milled to size and then extruded under strong pressure into pellets, a fast process that is highly suitable to medium- or largescale production. During the mechanical process, the raw

material is compacted approximately 3.7 times [1]. Additionally, the efficiency of burning wood pellets in small-scale combustion facilities ranges from 78 to 81%, making them the most effective instrument for bioenergy production on the small scale [1]. The Spanish wood-pellet market is to date small, but it is constantly increasing. For example, production of densified products quantified about of 50000 to 60000 tonnes, in 2004 [3]. However, in the last years, an important rise in compacted biomass fuels for residential purpose has taken place and pellet consumption is still mainly focused towards the small scale energy market (home heating) and consumption is currently increasing. In the use of biomass is generally necessary to carry out a series of previous operations of physical transformation and pretreatment. These operations make possible to obtain, also, products of high value-added, promoting amplification of their market and consumption. Nevertheless, in order to guarantee feasible previous conditioning, it is necessary that the increment of final product value compensates the material transformation cost. This way, as function of the final application, it will be necessary to apply different transformation stages. The main transformation stages for the pretreatment of the residual biomass are chipping, drying, milling, classification and densification. Among these, drying and milling are common to practically all the types of biomass or energy-using processes in some of their phases. Therefore, if this pretreatment is for cofiring, briquetting or pelletizing, woody biomass has some final conditions imposed by its final purpose; which implies different drying conditions. The detailed study of drying energy requirements is essential to guarantee the feasibility and efficiency of the process. Natural drying has many economical advantages but is not usually possible to decrease moisture content below 20% wet basis (w.b.), not even with appropriate climate and adequate infrastructure for the long-time storage. Because of that, if moisture content reduction achieved with natural drying is not adequate for the processing of the material, or the necessary conditions for their implementation are not available, forced drying may become the suitable alternative.

16th European Conference & Exhibition on Biomass for Energy, Biomass Resources, 2-6 June 2008, Valencia, Spain

Rotating drum-type (trommel) and pneumatic-type are the most widely-used equipment for biomass forced drying. Both are frequently operated with air indirect- or flue gas direct-heating, although in general, for the same temperatures there is no industrial explanation for the first one, unless the material should meet hygienic conditions. The trommel-type rotary dryer is often used for working with very wet materials and/or large particle sizes. In essence, it consists on the movement of solids across a hot fluid flow stream, which removes water of the material during its passing. In this apparatus, the flow channel is a rotating cylinder, which enables contact between solids and the drying medium, as shown in Fig. 1. In this type of dryer the movement of solids is caused by slope of the shell, cylinder rotation and the geometry of its inner surface.

Combustion chamber Fan air

wet material hot gas

humid gas exhaust to ambient

Dryer

natural gas

dry product

Figure 1: Layout and operation process of a conventional rotary dryer. a: Co-current rotary dryer; b: process. According to the heat flow direction, there are cocurrent and counter-current rotary dryers. In the first type, the flow of gas and solids goes in the same direction, therefore the gas increases humidity during its path and it is possible to become saturated with water at end, decreasing drying efficiency. In counter-current rotary dryer, flow of gas and solids travel in opposite directions. Thus, solids move across an increasingly dried environment with higher temperatures, bringing dehydration more effective and consistent than the first. However, the first one is more appropriate for biofuels, since presents lower ignition and firing risk. Typical inlet gas temperatures are usually between 200 and 500C and outlet between 90 and 120C (to prevent condensation and heat loss in the stack). Also, it is common to install a dust recovery system before the stack. A sketch of a typical co-current dryer system is shown in Fig. 1b. Rotary dryers have been specially and widely used in the mining, thermoplastic, fertilizers and foodstuffs industry. Since its origins in the last century, rotary dryers have been very thermally inefficient due to its poor design. Research in this area has been mainly aimed toward fluid bed or pneumatic dryers, even leaving aside the rotary one [1]. Due to the extensive use of this type of dryers for other applications, recent studies have been focused at the development of semi-empirical models to improve their design, operation and performance.

However, studies about different energy uses (i.e. biomass drying for co-firing or densified products) are, up to now, scarce. Apart from equipment sizing calculations, another relevant aspect is the scaling-up of energy consumption. Thus, studies of this type in dryers produce results depending on the nature of dried biomass, its composition and hygroscopic level. For example, required energy to remove 1 kg of water from a typical biomass fuel can exceed 2.6 MJ/kg of water, where typical LHV (lower heating value) between 18 and 21 MJ/kg of biomass, and moisture content of fresh, raw biomass like wood usually exceeds 50% wet basis [4]. Other literature about rotary dryer performance indicates that the required heat to evaporate 1 kg of water from wood chips is equivalent to 3.1 MJ [5, 6]. However, these data can change depending on biomass type and homogeneity of the material. For instance, moisture content and yield of biomass crops usually depend on climate conditions and harvesting period. Therefore, the homogeneity of biomass material cannot always be constant [7]. For co-firing purposes, there are no universal rules concerning the drying process economic feasibility of raw biomass. Each situation needs to be independently studied by considering the drying advantages and disadvantages (e.g., fuel saving, reduction of flue gas volume flow, higher combustion efficiency, more investment, etc.) [7]. On the other hand, drying is rather necessary for production of pellets or briquettes and a special pretreatment is key requisite to preserve product quality. Due to biomass drying process, it is necessary for each material to study in depth the relationships between pretreatment energy requirements compared with moisture content for each biomass size and type. Moreover, the knowledge about behavior of rotary drying equipment under different operation conditions can be also used to predict the moisture or temperature profiles of the solid and gas streams along the dryer, which are especially useful when the solid final quality is affected by the temperature levels reached inside the cylinder, as it is the case of densified products. This study is divided in two phases: an initial phase corresponding to this paper, in which initial energy requirements of woody biomass drying process are determined. Although other aspect as adaptation of installation to operate with forest residue was important, however it is not the purpose of this study. Lastly, a later phase in which the influence of initial moisture content, material size and type will be analyzed and energy consumption determined.

2 EXPERIMENTAL EQUIPMENT AND PROCEDURE 2.1 Description of equipment The semi-experimental rotary dryer used in this work is 4.5 m in length and 0.6 m in diameter. It consists of a rotating cylinder, which is slightly inclined to the horizontal to promote material flow. The rotary dryer is of co-current type due to the high biomass volatile content for which the control of final solids temperature was highly necessary. The inside of cylinder is fitted with material conveying flights and lifting flights of typical various forms designed to lift and drop the material

16th European Conference & Exhibition on Biomass for Energy, Biomass Resources, 2-6 June 2008, Valencia, Spain

through the gas stream as both material and gases move through the shell, thus enhancing close gas-solids contact; see details of flights in Fig. 2. This special flights design has been seldom studied in literature.

Table I: Rotary dryer design data. Item Dryer size (m x m) Max. evaporation (kg/h) Max. feeding flow rate (kg/h) Max. product humidity (%) Particle Size (mm) Rotating speed (rpm) Slope (m/m) Material Data 0.60 x 4.5 70 150 8 5-50 4-7 0.8 Wood chips

Table II: Instrument specifications. Measurement Instrument Accuracy Temperature k type thermocouple 1.5 C Differential pressure DP transmitter 2% of v. Humidity H/T transmitter 2% of v. Solids mass Weighing balance 0.5 kg Air velocity Pitot tube/DP transmitter 2% of v. Electrical power Power analyzer 0.5% of v. Moisture content Weighing balance 2% of v. Special instruments: in-drum, wireless H/T sensors traveling with solids, and material sampling fittings. The performance capacity of a continuous direct-heat rotary dryer must be demonstrated only under conditions of steady-state flow of material and gas. For steady-state conditions, the feed material rate, moisture content and temperature, air velocity, temperatures, humidity before and after of the drum, and product rate were remained essentially constant during the test period. An agreement within 10% of each data was assumed correct to determine steady-state without explanation [8]. According to this criterion, four of seven runs were valid for determining mass and energy balances. In the run AR-7 the steady-state condition was reached with two different air temperature (AR-7a and AR-7b), therefore five data were obtained in this phase of tests. In these AR runs, the solid product was discharged even wet (8.523.2% wb). The inlet wet-bulb temperature of the air was maintained at approximately the material feeding temperature of dryer. It is known that if biomass is very wet and dried products too, solid temperature inside of cylinder is substantially constant along the entire dryer length [9, 10]. This constant temperature was within the limits of the measurements, very proximity to the wetbulb temperature of the air in the dryer, except for a short time at very beginning of the dryer. These tests covered only a material, woody chip, ranging in average particle sizes from 5 to 50 mm, feed moisture content varied from 46 to 55% wb, material feeding rates from 132.6 to 168.2 kg/h, inlet air temperatures from 170.5 to 223.7C, air rates from 1674 to 1870 kg/h, and dryer holdups from 20 to 30%. Extra parameters like the inclination of the cylinder in 0,8 m/m, rotational speed in 4 rev./min., and the configuration of the flights were kept constant in all the runs (see parameters of runs in Table III). For each runs, about 600 kg of woody chips was loaded in hopper while air preheater was started and research facility prepared. In this work, control of drying process was done by means of varying air temperature via fuel input to the air preheater, while amount of inlet air was fixed in each run. It took up to an hour the initial

Figure 2: Intertwined flights inside of cylinder. Wet biomass is fed from a hopper by means of a screw feeder and a rotary valve which assure minimum air leakage. The feeding duct was specially designed to prevent plugs and backward movement. The dry product discharge is coupled to a rotary valve which seals the product exit and provides stable air flow. The operating configuration is co-current flow, with the hottest air in contact with the coldest product, because this limits product excessive heating, which is convenient for high volatile content of biomass. The ends of rotating cylinder are joined to stationary discharge ducts that connect to the air supply and exit gas ducts, material feed, and product conveyors. The annular clearances between the ends of the rotating cylinder and discharge ducts are enclosed by friction rotary seals in order to minimize the effect of air leakage on the operating conditions. The air stream is heated indirectly by an air preheater before entering the dryer shell. The air stream provides all the thermal energy needed to (a) heat the material and the moisture to be removed, heat and evaporate the moisture, and heat the vapors to exhaust temperature; and (b) compensate for conduction, convection, and radiation heat losses from the cylinder, discharge ducts, and ductwork, which most of them are isolated. After leaving the dryer cylinder, particle-laden hot humid air is filtered through a cyclone separator unit before being released to the atmosphere. The fan arrangement used on this dryer with low pressure drop air heater, is a double, forced- and induced-draft fan. In this manner, by balancing the two fans, the pressure inside the drum at the rotating seals can be maintained at a level close to, but slightly below, atmospheric pressure, i.e., between -0.2 and -1.0 kPa. This ensures that no process vapors leak to the atmosphere. Design specifications are shown in Table I. 2.2 Experimental runs The apparatus is equipped with standard instruments for measuring air flow, temperature and humidity. Initial and final solids moisture contents are measured by means of laboratory stove and weighing balance according to standards. In addition, in second phase of runs special instruments will be used. Details of instrumentation used are shown in Table II. The measuring errors of the main measurements categories are also given in the table.

16th European Conference & Exhibition on Biomass for Energy, Biomass Resources, 2-6 June 2008, Valencia, Spain

Table III. Results of drying runs.

Run No. AR-2 AR-4 AR-6 AR-7a AR-7b

Feed Product Air Feed Product Inlet Air Flow Flow Flow Moisture Moisture Product Wet-bulb Inlet Air Exit Air Rate, Rate, Rate, Content, Content, Temperature, Temperatures, Temperature, Temperature, kg/h kg/h kg/h % wb % wb C C C C 132.6 159.9 144.2 162.6 162.6 62.7 1674 46 10.8 44.3 51.4 181.2 85.9 1675 55 23.2 35 53.9 181.2 71.9 1870 49.7 8.5 51.5 58.5 218.9 70.1 1838 52.1 9.3 49 60.5 222.4 67 1729 53.6 13.7 39 60.5 223.7 Rotational Speed = 4 rev./min., Slope = 0,8 m/m, Material: Woody chips. 81.6 71.5 103 100.6 99.5

heating, meanwhile depression in cylinder was set. Wood bags were weighed prior to use. Discharged products were weighed in continuous to verify steady flow. All measurements were recorded via SCADA stands for control and data acquisition, at minimums intervals. 2.3 Evaluation parameter Literature data were used to compare with this study. Comparison of drying energy requirements is based on the following performance indicators. Specific energy utilization (SEU) was defined as sum of electrical and fuel energy used per unit mass of dried solids produced [10]: (MJ/kgdry solids), and (1) Table IV: Thermal energy balance results. thermal drying efficiency was defined as the heat used to evaporate moisture from the woody biomass by the heat input from fuel [10]: (2) Thermal energy utilization (MJ/kgdry solids) 5.7 2.7 2.6 0.2 0.1 0.1 Percent of total (%) 100% 48% 47% 3% 1% 1%

Figure 4: Variation of relative humidity during drying.

Item Energy input Heat for drying Flue gas loss Radiation loss Leakage loss Other losses

RESULTS AND DISCUSSION

Figs. 3 and 4 show typical variations of air temperature and relative humidity at inlet and outlet of cylinder as a function of drying time in comparison with ambient values. The air temperature was maintained between 210 and 220C during this run AR-6. It was spent approximately 60 minutes to obtain uniform products discharge and stability conditions observed. It was necessary several hours of previous drying to end up knowing the installation and to obtain consistent results. Table IV presents thermal energy balance of semiexperimental rotary dryer per kgdry solids.

As it is appreciated in Table IV, thermal energy is only considered for energy balance; this is because power energy required moving ancillary equipment was 5% from the thermal one. Therefore, if the whole energy is taken into account, SEU specific energy utilization was found to be 6.0 MJ/kgdry solids (0.3 MJ/kgdry solids for electricity), which are required to drive ancillary equipment (screw feeder, rotary valves, fans and cylinder driver systems). Respect to thermal drying efficiency, just as it was previously defined, the results obtained in this study pointed to that the heat used to evaporate moisture from material was 46% from heat input from fuel, this is 2.5 MJ/kgwater. However, typical rotary dryers loss heat by inleaked air heating and external radiation. Taking into account this unavoidable energy consumption, the required heat ascended up to 2.8 MJ/kgwater. The value observed was between those derived from [4, 6] with 2.6 and 3.1 MJ/kgwater respectively. It should be noted that [4] reported for any biomass fuel, not necessary woody chips, as it is the case.

Figure 3: Temperature evolution during drying.

16th European Conference & Exhibition on Biomass for Energy, Biomass Resources, 2-6 June 2008, Valencia, Spain

CONCLUSIONS

Dry woody biomass as energy source is a very important renewable energy resource for Spain and other countries of OECD. Improvement of energy efficiency in the drying process will help to profit its use in power generation and accomplished CO2 reduction targets. In this paper, mass and energy balances in a semiindustrial rotary dryer have been carried in order to evaluate the specific energy utilization and thermal efficiency of woody biomass drying. On basis of experimental results from this study compared to literature data, it has been established that it is necessary about 2.7 MJ/kgdry solids and 2.8 MJ/kgwater to evaporate 1 kg of water from around 50 to 10% wb. Ancillary equipment, e.g., screw feeder, rotary valves, fans and cylinder driver systems, consume up to 5% of heat for drying per kg dry solids with which overall energy increase to 3.0 MJ/kgdry solids. In an experimental second phase, is planned to perform tests including dedicated instrumentation: wireless H/T solid-travelling sensors and material sampling fittings will be necessary to obtain air temperature profile and biomass moisture content at different drum axial locations. Moreover, it will be necessary in the next future to analyze how to scale-up energy requirements that was obtained in this work, using different types and size of biomass.

ACKNOWLEDGEMENTS

This study is part of the projects ENE 200405137/ALT, funded by the Spanish Ministry of Education and Science and the project CENIT CO2 (DT 02 09 SA OA) funded by CDTI. We wish to thank CESEFOR and Unin Fenosa Generacin for supplying the forest residue biomass.

[3] Passalacqua, F., Zaetta, C. Pellets in Southern Europe. The state of the art of pellets utilization in Southern Europe. New perspectives of pellets from agri-residues.. Rome, Italy : s.n., 2004. [4] Svoboda K., Pohorely M., Martinec J., Baxter D., Hunter Ch., Integration of biomass drying with combustion/gasification technologies and minimalization of organic compounds, 32nd International Conference of SSCHE, Slovakia, May 23-27, 2005. [5] Williams-Gardner A. Industrial Drying. Leonard Hill Books, 1971. [6] Brammer J., Bridgwater A. Drying technologies for an integrated gasification bio-energy plant. Renewable and Sustainable Energy Reviews 3 (1999) 243-289. [7] Maciejewska, A., Veringa, H., Sanders, J., Peteves, C. Co-firing of biomass with coal: constraints and role of biomass pre-treatment. DG JRC and Institute for Energy. 2006. [8] AIchE Equipment Testing Procedure. Continuous Direct-Heat Rotary Dryers: A guide to performance evaluation, 3th ed. Chicago: Wiley, 2004. [9] Friedman, S. J. y Marshall W. R., Studies in rotary drying Part II heat and mass transfer. Chem. Eng. Progr., 45, 482, 573 (1949). [10] Tippayawong, N., Tantakitti, S., Thavornun, S. Energy efficiency improvements in longan drying practice. Energy. (2008). [11] Mujumdar AS, editor. Handbook of Industrial Drying, 1. Marcel Dekker, 1995. [12] Perry, R.H. y D.W. Green (eds). Perrys Chemical Engineers Handbook, 7a. ed. Nueva York: McGrawHill, 1997: pp. 12. [13] McCabe, W, Smith, J y Harriot, P. Operaciones Unitarias en Ingeniera Qumica, 6. ed. Mexico: McGraw-Hill, 2001.

NOMENCLATURE

AR CHP OECD

Adiabatic drying run series Combined heat and power stations Organisation for Economic Co-operation and Development LHV Lower heating value SCADA Stands for control and data acquisition SEU Specific energy utilization wb wet basis Energy of fuel Energy of electricity Mass of dry solids Mass of water

REFERENCES

[1] Mabee, W., Saddler, J. Forest and Energy Working Paper: Forests and energy in OECD contries. s.l. : Food and Agriculture Organization of the United Nations, 2007. [2] Instituto para la Diversificacin y el Ahorro de la Energa de Espaa. Annual Report 2006.

Potrebbero piacerti anche