You are on page 1of 50

EPSRC Thermal Management of

Industrial Processes





A Review of Drying Technologies

(February 2010)




Report Prepared by:
SUWIC, Sheffield University

Researchers: Dr Hanning Li, Dr K Finney

Investigators: Professor Jim Swithenbank
Professor Vida N Sharifi




Sheffield University Waste Incineration Centre (SUWIC)
Department of Chemical and Process Engineering
Sheffield University




Executive Summary

In accordance with the EPSRC grant proposal, Sheffield University has conducted an extensive
literature review on biomass drying and an evaluation of the drying process. This report presents
the results obtained from our literature review work. Various sources of information were used in
order to compile this report. These included websites, journal publications, reports and
communications with manufacturers and industry.

The main topics covered in this review work include:

Biomass Drying (benefits and drawbacks)
Drying equipment and processes
Assessment of drying technologies when using flue gases
Assessment of drying technologies when using superheated steam
Costs and environmental impacts of drying biomass
Recent technological developments when using low temperature heat sources for drying
biomass


This report present the findings from the above review work.




Acknowledgements:
The authors would like to thank the Engineering and Physical Science Research Council (EPSRC
Thermal Management of Industrial Processes Consortium) for their financial and technical support
for this research work.






List of Contents
1. Introduction.................................................................................................................................. 4
2. Classification and Properties of Biofuels..................................................................................... 4
2.1 Benefits of Drying Biomass for Combustion and Gasification .......................................... 7
2.2 Drawback of Dried Fuel...................................................................................................... 8
2.3 Background of Biomass Drying........................................................................................ 10
3. Drying Process ........................................................................................................................... 11
3.1 Stages of Drying................................................................................................................ 12
4. Drying Equipment and Processes .............................................................................................. 13
4.1 Rotary Dryer ..................................................................................................................... 13
4.2 Flash Dryer........................................................................................................................ 16
4.3 Fluidized Bed Dryer.......................................................................................................... 18
4.4 Sprout Bed Dryer .............................................................................................................. 19
4.5 Belt Dryer.......................................................................................................................... 20
5. Assessment of Dryer Technologies using Flue Gas ................................................................... 21
5.1 Performance ...................................................................................................................... 21
5.2 Heat Recovery................................................................................................................... 23
5.3 Fire Safety in Dryers ......................................................................................................... 25
5.4 Environmental Aspects ..................................................................................................... 25
5.5 Cost ................................................................................................................................... 27
5.6 Dryer Selection ................................................................................................................. 29
6. Superheated Steam Systems....................................................................................................... 31
6.1 Dryer Type ........................................................................................................................ 31
6.2 Advantages and Disadvantages......................................................................................... 33
6.3 Capacities of Dryers.......................................................................................................... 35
6.4 Performances..................................................................................................................... 35
6.5 Heat Recovery................................................................................................................... 36
6.6 Cost ................................................................................................................................... 37
6.7 Environmental Aspects ..................................................................................................... 38
7. Recent Development in Low Temperature Heat Sources for Biomass Drying.......................... 39
7.1 DRY-REX......................................................................................................................... 40
7.2 SRE- Renergi LTD Dryer.................................................................................................. 43
7.3 Microwave ........................................................................................................................ 45
8. Conclusions................................................................................................................................ 46
9. References.................................................................................................................................. 48


4

1. Introduction
Dry biomass fuel provides significant benefits to combustion and gasification, but they must be
balanced against increased capital and operation costs. Currently, several methods have been
established and some promising technologies are being investigated.
2. Classification and Properties of Biomass
Material of biological origin, excluding material embedded in geological formations and
transformed to fossil fuels, is called biomass (Holmberg, 2007). As a renewable energy source,
biomass is derived from living organisms, such as wood, herbaceous crops and waste. Woody
biomass is composed of bark, forest residues, sawdust and cutter shavings. Herbaceous biomass
is grown from numerous types of plants, including miscanthus, switch grass, hemp, corn, poplar,
willow, sorghum and sugarcane. Biomass waste includes construction wood, crushed or chipped
used wood, used paper, pulp and paper sludge, municipal solid waste (MSW), manufacturing
waste and landfill gas. Figure 2-1 shows some samples of solid biomasses.

Biomass is carbon based matter and is composed of a mixture of organic molecules containing
hydrogen, oxygen, nitrogen and also small quantities of other atoms, including alkali, alkaline
earth and heavy metals. Table 2-1 presents the properties of biomass fuels.
5


Bark Wood chips Pulp and paper sludge

Crushed construction wood Sawdust Sugarcane bagasse
Figure 2-1. Examples of biomass materials.



6

Table 2-1. Typical physical and thermochemical properties of selected wet biomass fuels (Bruce and Sinclair, 1996).
Pulp and Wood Residues Milled Lignite Sugarcane

Paper
sludge Bark
Fuel
Chips Peat Bagasse
Moisture % wet basis 50 - 65 30 - 60 45 - 55 45 - 55 30 - 50 48 - 52
HHV, dry basis MJ/kg 15 - 19 19 - 25 19 - 21 19 - 21 16.6-24.3 18.6-20.3
HHV, wet basis MJ/kg 6 @35% 11@50% 10@50% 10@50% 12@40% 10@50%
Bulk Density (wet
basis) kg/m3 500-900 290-380 260-320 300-400 650-780 80-130
Volatiles % dry wt - 69 - 76 70 - 85 60 - 70 50 - 60 84

Ultimate analysis
Carbon % dry wt 25 - 50 55 50.6 50 - 60 41.5-61.4 43.2-49.0
Hydrogen % dry wt 3 - 6 5.8 6.2 5.0 - 6.5 3.4 - 4.6 5.9 - 6.6
Oxygen % dry wt 19 - 38 39 43 30 - 40 11.3-19.7 43.4-48.0
Nitrogen % dry wt 2 - 5 0.1 0.1 1.0 - 2.5 0.7 - 1.1 0.1 - 0.41
Sulphur % dry wt 2 - 48 0.1 0.02 0.1 - 0.2 1.0 - 2.4 trace
Ash % dry wt 3 - 50 3 0.1 2 - 10 11.9-40.6 1.4 - 2.9
7


2.1 Benefits of Drying Biomass for Combustion and Gasification
It is commonly accepted that wet fuel consumes some heat of combustion to evaporate water in the
fuel. As shown in Table 2-1, the average moisture content of biomass is typically between
55-60% (wet basis) depending on the biomass type. In addition to the high average moisture
content, the weather, season and storage time may cause drastic deviations in the bark moisture.
The moisture content of sawdust may vary from air-dry to 70% wet basis (Holmberg, 2007). The
high moisture in the biomass will have a negative effect on combustion and gasification operations.
The water should be removed from the biomass to improve the quality of combustion and
gasification.

The heating value (i.e. HHV, as shown in Table 2-1) is dependent on biomass moisture, where a
lower moisture content in the biomass increases the heating value. As a result of drying, energy
input into the boiler may be increased without increasing the fuel input, or the fuel input into the
boiler may be decreased to get the same energy input, as in the case of moist fuel. The basic
functions of the combustion control and burner management systems are to maintain constant
steam flow or pressure under varying loads, through proper input of fuel and to maintain safe and
efficient operation throughout the boilers load range. Compared with several other fuels (e.g. oil,
natural gas and coal), the heating value of biomass varies as a result of varying moisture content.
Generally, it is possible to control large changes in fuel quality during the combustion, but such
boilers set high requirements for the process control. Drier biomass will significantly reduce the
range of varying water content, facilitating process controls in combustion processes.

With dry fuel, all the heat of combustion goes into heating the air and products, leading to a flame
temperature of 2300-2500F (1260-1370C), while green wood has a combustion temperature of
about 1800F (982C) (FBT, Inc., 1994). The increased flame temperature is beneficial in many
aspects. First, the higher flame temperature means that there is a larger temperature gradient in
the boiler for radiant heat transfer. More heat transfer takes place for the same boiler tube area,
increasing steam production. In new boilers designed for dried fuel, the boiler can be smaller
because less heat transfer area is needed.

The high flame temperature with dry fuel could achieve more complete combustion of the fuel,
8

resulting in lower carbon monoxide (CO) levels and less flyash leaving the boiler. More
complete combustion also means more heat is released from the fuel. In a new boiler, the fire
box and the downstream ash handing system can be smaller.

With better combustion, the extra air can be reduced. This reduction in excess air means less heat
of combustion goes into heating air. Using less excess air also reduces sensible heat losses with
the flue gases, hence increasing the boiler efficiency. The forced draft (FD) fan, which provides
air for combustion, will consume less power with less excess air. Likewise, the induced draft (ID)
fan, which draws the flue gas out of the boiler and through the pollution control equipment, will
require less power.

There will be an increase of up to 5-15% in the overall thermal efficiency, and possibly up to a
50-60% increase in steam production (Wade, 1998).

2.2 Drawback of Dried Biomass Fuel
Drying biomass is expensive and the additional costs may discourage the use of dried biomass fuel.
Figure 2-2 presents the costs for wood (ELECTROWATT-EKONO, 2003). The total production
cost of pellets ranges from 52.2 to 81.3 /t (without drying) and from 73.5 to 94.6 /t (with drying).
The drying costs are dependent on the technology used and range from 25 to 29 /t
pellets.
The
production costs are also dependent on the annual operational time. If a plant is run in a 3 shifts,
7 days a week mode, the production costs are approx 84 /t
pellets.
Other operational modes and the
associated costs include: 3 shifts and 5 days a week = 90 /t
pellets,
2 shifts and 5 days a week = 101
/t
pellets
and 1 shift and 5 days a week = 133 /t
pellets.

The drying is an additional operational cost in the process and power industries, even though this
may be offset by using smaller boilers, air emissions equipment and fuel handling equipment. If
the required biomass is over dried, the energy consumption and maintenance costs could be
significantly increased. The additional complexity may also affect the overall system operation.

9


Without drying With drying
Figure 2-2 Pellet costs without/with drying

In addition to the cost concerns, there are other operational and environmental issues, which must
be addressed when using wet biomass material.

Ash deposition on heat exchanger tubes is commonly known to reduce the heat transfer, which will
subsequently result in a reduction in plant thermal efficiency. Ash deposition on heat exchanger
tubes is also a precursor for corrosion and fouling processes in heat exchangers. Shao, et al.
(2009) found that the dried biomass increased the ash deposition rate during biomass combustion.
Such ash deposits from dried biomass decreases the amount of biomass available for the
combustion and increases the maintenance costs (i.e. plant downtime to remove deposits).

Burning dried biomass fuel results in higher combustion temperatures in the boiler. However, as
the flame temperature increases, it approaches the fusion temperature of the ash. If the ash starts
to flow and form slag, this can seriously affect the boiler operation. Usually the flowing
temperature of the ash is safely above the flame temperature, but when contaminants from
construction debris or salts are mixed with the fuel, the flowing temperature can be lower.

The major problems of slagging are associated with alkali metal content, principally sodium and
potassium, both common in biomass fuels with high ratio of alkaline metal oxides to silica. As
dried fuel is burnt in the furnace, the increased ash leads to high concentrations of sodium and
potassium and thus a tendency to slagging at lower temperatures.
10


Slagging can seriously affect the boiler operation. As a consequence, heat transfer tube life-time
is reduced and maintenance costs are increased. New material can be designed for corrosion
resistance, but the cost is a concern.

With drier fuel, the higher furnace temperatures will tend to increase the formation of nitrogen
oxides, though modern over fire air systems could minimise NOx concentration (compared to
older designs). When NOx emissions from the firing of biomass fuel (even after drying), are
compared with the natural gas-fired conventional burners, the levels are quite similar. However
NOx emissions from biomass fired plants are significantly lower when compared with the oil and
coal fired systems. Low NOx-burners are an active area of development and some natural gas and
oil burners are capable of very low emissions, significantly lower than that from a conventionally
designed biomass boiler.

2.3 Background of Biomass Drying
Even though biomass drying is not common before combustion at the moment, commercial
biomass dryers were used at pulp and paper mills in the 1970s and 1980s. The dominant
combustion technique for biofuels at that time was grate firing. This type of boiler can handle
fuels with varying moisture contents, but ideally a moisture content of 30-40% wet basis should be
used (Wimmerstedt 1995). The main reason for investments in manufacturing and designing
dryers in the 1970s and 80s was probably the high oil prices resulting from two oil crises. In
some cases, the moist fuel also decreased the boiler capacity so much that it became reasonable to
install a dryer in combination with the boiler. In the 1970s and 1980s, industrial dryers were
direct flue gas dryers (Holmberg, 2007). Flue gases were either taken directly from the boiler or
generated in a separate flue gas burner. The most common dryer types were drum dryers and
flash dryers.

Since the 1970s, fluidized bed boilers have replaced grate firing as a combustion technique
(Huhtinen, 1999). Compared with grate firing, the fluidized bed boiler is a more suitable
combustion technique for moist biofuels. However, the use of moist fuel decreases the energy
efficiency of the power plant. One solution could integrate the existing flue gas dryers with
fluidized bed boilers. Bad operating experiences, environmental considerations and economic
11

factors may block the integration development.

Biomass is a burnable material with a heterogeneous particle size. It is also typical that the
biomass flow contains stones and sand. It is obvious that the properties of biofuel set tough
requirements for the operation of the dryer. For some dryer types (e.g. flash dryers), the
maintenance cost may also be high. It is presumable that the operational experiences of the flue
gas dryers have not always been satisfactory and bad experiences have supported the decision to
develop the dryer.

All types of wood material contain volatile organic material that may be emitted together with the
water vapour, as shown in Table 2-1. The emissions of biomass drying are greatly affected by the
drying temperature, when it exceeds 100C (Holmberg, 2007). Below 100C, emissions are
reported to be low (Spets, 2004). The drying temperatures of flue gas dryers are clearly higher
than 100C. Exhaust gases or unclean condensates must be treated after the dryer if they contain
high concentration of emissions. Treatment increases drying costs and the treatment of the
exhaust gas may be a cost issue.

3. Drying Process
Biomass drying systems consist of three principal factors: drying medium, heat supply and dryer
type. Drying medium are mostly flue gas, air and steam. Heat supplies into the biomass are
operated through convection (direct dryers), conduction (indirect dryers) or the combination of
direct and indirect dryers. The type of dryers includes rotary dryers, flash dryers, belt dryers and
fluidized bed dryers, among others. Rotary and flash dryers are mostly used in industry today.

In direct-heated dryers, hot air, flue gas or superheated steam is in contact with the biomass
material. The hot air, flue gas or superheated steam loses its sensible heat and provides the latent
heat of evaporation to dry the materials. The air also removes the water vapour. The material
can be agitated by mechanical devices or by fluidized air. The super-heated steam remains above
its saturation without condensation.

Indirect drying separates the biomass material from the heat source (either hot air or superheated
steam) by a heat exchange surface. As a result, the latent heat of evaporation of the water vapour
12

is easy to recover since the water vapour is not diluted by air. With super-heated steam drying,
the dryer can be designed to produce steam at a desired pressure.
3.1 Stages of Drying
There are several steps for drying. First, the material must be heated up to the wet bulb
temperature to produce a driving force for water to leave the wet material. Next, any surface
moisture is evaporated very quickly. Once all the surface moisture is removed, the material must
be heated to drive water from the inside of biomass to the surface so it can evaporate. This
happens when the rate of drying drops as the material remains close to the wet bulb temperature.
Once the material is completely dry, it begins to heat up to the surrounding temperature, because
water is no longer present to keep the temperature low.

These steps are important for drying a combustible material. High temperatures are desirable to
increase heat transfer and minimize equipment size; but at same time, the fuel ignition could be a
concern. By understanding these steps involved with biomass drying, fast drying at high
temperatures can be exploited with minimal fire risk.

Significant fire risks mostly occur in two instances. The first is after the surface moisture has
been evaporated, but before an appreciable amount of water has been driven out from inside the
biomass. During this very short period, no water vapour at the surface keeps the fuel particle
cool, leading to its surface quickly heating up while the inside remains cool. If the surface
remains hot for a long enough time, the biomass can ignite even if it is not completely dry.
However, once the inside of the biomass starts to drive water to the surface, this supply of
moisture will keep it cool until it is completely dry (Intercontinental Engineering, Ltd. 1980).
Over-dried biomass is also a fire risk. If the biomass loses all its moisture, it will begin to warm
and can ignite when it reaches its combustion temperature. Generally speaking, dryers are not
designed to completely dry material.

When a material still has moisture associated with it, its temperature will be very close to the wet
bulb temperature of air as evaporation occurs, regardless of air temperature. Therefore, a very
hot air stream can be used to dry biomass in a co-current flow process because the hot air is
introduced to the dryer along with the wet biomass.

13

In super-heated drying, there is no wet-bulb temperature because only steam is present. The
water in the fuel must instead be heated to its saturation temperature before it evaporates from the
biomass, but once it turns to vapour, it does not need to diffuse through the air to get out of
biomass or have saturated air removed from the surface to promote evaporation. As long as
material temperature is higher than the saturation temperature, the vapour pressure of the water
will cause the moisture to flow out of the material. In superheated process, the material will stay
at its saturation temperature until it is completely dry, then the temperature will start to increase,
the same as the air-dried case.

Because no oxygen is present in super-heated steam drying, the fuel cannot burn, even at elevated
temperatures. However, there is one potential risk of fire, i.e if the material is allowed to dry
completely and heats up above its ignition temperature of about 260C.

Despite the medium used in the drying process, potential fuel ignition should be a concern in the
equipment and process design, as well as the selection of dryers.

4. Drying Equipment and Processes
Currently, various dryers are used in industry. The following descriptions introduce those that are
mostly used in industry, including rotary, flash, belt, sprout bed and fluidized bed dryers.
4.1 Rotary Dryer
Rotary dryers are the most common type for biomass drying. Despite the introduction of new
technologies, the long-applied rotary dryer is still widely regarded as the workhorse of the drying
process industries. The robust yet simple construction combines flexibility with reliability,
enabling this type of dryer to handle a vast range of materials and to operate continuously under
the most arduous conditions. The design also permits the use of the highest possible drying
temperatures and, in contrast to other dryers, is not sensitive to wide variations in material size,
moisture content or throughput.

In its simplest form, the rotary dryer consists of a slightly inclined rotating cylinder, fitted with a
series of peripheral flights arranged to lift, distribute and transport the material. The flights are
designed to suit the particular handling characteristics of the material, which may vary with
14

increasing dryness. Figure 4-1 shows internal structure of flights in a typical rotary dryer made
by Aeroglide Corporation and GEA Barr-Rosin .


Figure 4-1. Internal construction of a rotary dryer made by Aeroglide Co. and GEA Barr-Rosin
http://www.aeroglide.com/rotary-dryers.php; http://www.barr-rosin.com/products/products.asp

There are several types of rotary dryers, but the most widely-used is the directly-heated single-pass
rotary dryer. In this type of dryer, hot air or gas is in contact with biomass material inside a
rotating drum to induce the evaporation of the moisture. The rotation of the drum, with the aid of
flights, lifts the solids in the dryer so they fall down through the hot gas, promoting better heat and
mass transfer. If contamination is not a concern, hot flue gas can also be fed directly into the
dryer. The heat in the hot air or gas evaporates the water and consequently, the gas temperature is
rapidly reduced as it leaves the dryer. The exhaust gases leaving the dryer may pass through a
cyclone, multi-cyclone, gashouse filter, scrubber or electrostatic precipitator to remove any fine
material entrained in the air. An ID fan may or may not be required depending on the dryer
configuration. If one is needed, it is usually placed after the emission control equipment to
reduce erosion of the fan. It may also be placed before the first cyclone to provide the pressure
drop through downstream equipment. Figure 4-2 is a schematic diagram of rotary dryer and its
15

associated processes. The biomass and hot air normally flow co-currently through the dryer, so
the hottest gases come in contact with the wettest material, as shown in Figure 4-2. For materials
where temperature is not a concern, the flue gas and solid flow in opposite directions, called
counter-current flow, so the driest solids are exposed to the hottest gases with lowest humidity.
This counter-current flow configuration produces the lowest moisture in the biomass as it leaves
the dryer, but this exposes essentially dry material to a high flue gas temperature, which could
increase the fire risk.

Figure 4-2. Schematic diagram of co-current operation in a rotary dryer.

The basic, single-pass rotary dryer design can be modified to allow three passes of the air and
biomass through the dryer. The material first enters an inner cylinder with the hot air. Smaller
and drier material is quickly blown through the cylinder into a larger concentric cylinder for the
second pass. Larger material is moved and tumbled with the aid of flights. After the second
pass, the air and material pass back up the outermost cylinder of the dryer and are removed. The
triple-pass design works best with biomass smaller than an inch, because larger material can cause
plugging. Single-pass dryers can take larger material.

Another commonly applied technology is indirectly heated rotary dryers, which use a heat source
steam or hot air passing through the outer wall of the dryer or through an inner central shaft to
heat the dryer by conduction. This is more common with materials that would be contaminated
by direct contact with flue gases or with materials that react with air. A hybrid direct/indirect
16

rotary dryer also exists where very hot flue gases enter the dryer through a central shaft and
initially provide heat indirectly by conduction, then the same gases pass through the dryer coming
into direct contact with the wet material. During the second pass, the indirect heating warms the
flue gas and solids. In this way, a high flue gas or burner temperature can be used for heating,
while reducing the fire risk by limiting the temperature of the gas in direct contact with biomass.

The inlet gas temperature to rotary biomass dryers can vary from 230 to 1100C and the outlet
temperatures vary from 70 to 110C. Most dryers have outlet temperatures higher than 100C to
prevent the condensation of acids and resins. Retention times in rotary dryers can be less than a
minute for small particles, and 10-30 minutes for larger materials (Intercontinental Engineering,
Ltd., 1980; Haapanen, et al., 1983).

The efficiency of the dryer is largely dependent on the differential temperatures between the inlet
and exhaust gas, although the heat transfer rate is also influenced by the relationship between the
design of flights and the speed of rotation. Irrespective of the gas and material temperatures,
however, the drying (or residence) time may be critical, as this is governed by the rate of diffusion
of the water from the core to the surface of the material.

Numerous applications of rotary dryers have been made in the drying of sludges from municipal
waste water treatment plants. Relative to other sludge drying processes, the dryer produces large
quantities of exhaust gases containing odorous compounds. Other applications include pulp and
paper, saw mills and board mills. The manufacturers supplying various rotary dryers include
M-E-C Inc, Rader Inc., Raytheon (Stearns-Roger Division) and ABB Raymond (Bartlett-Snow).
4.2 Flash Dryer
The pneumatic or flash dryer dries biomass rapidly owing to the easy removal of free moisture or
where any required diffusion to the surface occurs readily; drying takes place in a matter of
seconds. Wet material is mixed with a stream of heated air (or other gas), which conveys it through
a drying duct where high heat and mass transfer rates rapidly dry the product. Flash or pneumatic
dryers are followed by a cyclone. The gas passes through a scrubber to remove entrained
particulate material. A simplified process of a flash dryer (without a scrubber) is shown in Figure
4-3.

17

The flash dryer provides short drying times and the equipment is more compact than a rotary dryer.
The flash dryer requires smaller sizes of biomass so as to transport and suspend it by an air stream.
Some dry biomass can be mixed with the wet inlet material to improve material handing. Flash
dryers have been used not only for wood waste, but also for peat, bagasse, lignite and other solids.
Gas temperatures are slightly lower for flash dryers than for rotary dryers, but they still operate at
temperatures above the combustion point. The solids retention time in a flash dryer is generally
less than 30 seconds.

Manufacturers include Flakt, of Sweden, Raymond Division of ABB Raymond, Ahlstrom of
Finland, Williams Patent Crusher and Pulverizer Co., Inc. of St. Louis, MI, F.L. Smidth, of
Denmark (Figure 4-4).

Figure 4-3 flash dryer configuration
18


M.E.C. flash tube dryer Two flash dryers (15t/h evaporation. Bar-rosin)
Figure 4-4. Constructed flash tube dryer.

4.3 Fluidized Bed Dryer
In a fluidised bed dryer, as shown in Figure 4-5, the hot air is supplied to the bed through a special
perforated distributor plate and flows through the layer (or bed) of solids at a velocity sufficient to
support the weight of particles in a fluidised state. Bubbles form and break within the fluidised
bed of material promoting intense particle movement. In this state, the solids behave like a
free-flowing boiling liquid. As a result, the gas phase achieves good contact with the biomass
particles, leading to well mixed gases and solids. Thus, fluidized beds operate at high heat and
mass transfer rates to provide uniform and fast evaporation. In fluidized beds, the distributor
plate is required to allow fluidising gas to be evenly distributed across the area of the bed.

During processing with fluidized beds, many biomass materials may become a sticky or cohesive.
Those with vibrating equipment, however, could be effective in reducing agglomeration. In
addition, a rotating agitator can be incorporated within the first drying section of the bed. This
slow-moving device serves to gently agitate the wet material, encouraging even fluidisation and
eliminating rat-holing without causing particle degradation. The agitation also helps the gas
phase to break up into bubbles that will improve heat and mass transfer.

19

Fluidized bed dryers are mostly used in steam drying processes and co-location near process plants
or power plants. The manufacturers include GEA Barr-Rosin Inc, Canada; GEA Process
Engineering Inc; Anhydro.


Figure 4-5. Schematic diagram and manufactured fluidized bed dryers.
http://www.anhydro.com/content/us/products/dryers/fluid_bed_dryers
http://www.barr-rosin.com/products/fluid-bed-dryer.asp
4.4 Sprout Bed Dryer
Sprouted dryers (cascade dryers) are commonly used for dying grain, although they can be used
for other types of biomass. The material is introduced by a flowing stream or screw driver and
the drying media is introduced at the bottom (Figure 4-6). The feedstock then falls down to the
bottom and is lifted again. The material is let out through the opening holes on the side of
chamber. The residence time is generally a couple of minutes.

The original cascade dryer design was by Bahco of Sweden. The original Bahco group is now
part of the ABB group which have retained the rights for the rest of the world. Others include
Hercules, Canada and ESI Inc. of Kenneshaw, GA. The cascade dryer is mostly be used for
wood waste.
20


Figure 4-6. Sprouted dryer (Berghel et al., 2008).
4.5 Belt Dryer
In belt dryers (Figure 4-7), the feedstock is spread onto a moving perforated conveyor to dry the
material in a continuous process. Fans blow the drying medium through the conveyor and
feedstock. If multiple conveyors are used they can be in series or stacked (i.e. multi-pass).
Belt dryers are very versatile and can handle a wide range of materials. Recently, belt dryers are
frequently used in low temperature operations to save energy, as well as to reduce air emission and
fire hazards. Belt dryers are provided by Swiss Combi, Bruks Klckner, Mabarex, Andritz Fiber.


Figure 4-7 Drying plant in Nufri, Spain. Plant type: Belt drying plant. Input: Sludge from fruit
juice plant. Heat Generation: waste heat, hot water. Evaporation: 3.2-5.3 t/h. Start up: 2005

21

5. Assessment of Dryer Technologies using Flue Gas
Thermal drying systems for moist fuels can be grouped into two categories, flue gas/air type and
steam type. Flue gas/air dryers use the sensible heat in the products of combustion to accomplish
the drying and are a well established technology. The heat supplied in the gases, together with
the moisture driven off, is directed first to the emission control equipment and then to the stack for
release. Steam dryers use the latent heat in steam to accomplish the drying. The steam
originating from the moisture driven off from the fuel is recirculated as the heat transfer medium.
They are a more recent development, and come in a variety of forms. Their assessment will be
discussed in next section.

The following introduces the advantages, limitation, manufactures, capacity, operational
conditions and costs of each dryer under flue gas/air as drying medium.
5.1 Performance
Rotary Dryer
Rotary dryers are less sensitive to particle size and can accept the hot flue gases. They have low
maintenance costs. The material moisture is hard to control in rotary dryers because of the long
lag time (Fredrikson, 1984). Rotary dryers also present a fire hazard and require the most space.
Compared with single-pass dryers, triple-pass dryers have higher capital costs, higher maintenance
costs, higher blower costs and pose more of a fire risk (Intercontinential Engineering, Ltd. 1980).

The retention time for the smallest particles can be as short as 30s, while for most of the material it
is in the order of 20-30 minutes. Common design features affecting retention time include single
and triple pass, and dense versus open internal flighting. The single pass dryer is less prone to
plugging and can handle a wider range of particle sizes; a triple-pass dryer uses concentric
cylinders to increase the gas path length and generally cost less. Some designs recirculate
55-65% of the flue gas to lower the gas temperature at the inlet to the dryer and reduce the
generation of blue haze fine particulates composed of the more volatile components of the
material being dried, which is particularly problematic in the case of wood.

Rotary dryers are suitable for materials which are in the form of free flowing solids and fibrous
particles of up to 125mm (5"), granules, pellets, broken filter cakes and powders. Additionally,
22

these materials must be moderately insensitive to heat (Bruce and Sinclair, 1996; Wade, 1998).

Evaporation capacities are mostly in the range of 3-23t/h, corresponding to dryer sizes of 3 feet
diameter x 8 feet to 13 feet diameter x 60 feet long. Quebec, Canada, has a rotary dryer with an
evaporative capacity of 23 t/h of 18.5 feet diameter x 85 feet long for a board mill. The rotary
dryer performance is summarized in Table 5- 1.

Table 5-1. Typical range of design and performance data for various dryers.
Rotary Flash Sprouted Belt
Evaporation (t/h) 3-23 4.8-17 2-41 0.5-40
Temperature (C) 200-600 150-700 160-280 30-150
Capacity (t/h) 3-45 4.4-16 1.5-25
Feed moisture (%) 45-65 45-65 45-65 45-72
Moist. Discharge (%) 10-45 10-45 10-45 15-25
Typical feed discharge 55,12 55, 35 58, 25
Pressure drop (kPa) 2.5-3.7 7.5 2.5-3.7 0.5
Particle size (mm, opt) 19-50
Particle size (mm, max) 25-125 0.5-50 100
Thermal Requirement
(GJ/t_evap) 3.0-4.0 2.7-2.8 2.7 1.3-2.5

Flash Dryer
Flash dryers are much more compact than rotary dryers, but have higher installation costs
(Fredrikson, 1984). They can be used on most types of biomass, but have high blower power
costs in addition to the heat requirements for drying. The particles being dried must be small
enough to be suspended in the air stream. Heat recovery is difficult due to the fact that the air
mixed with the water vapour. Flash dryers have a lower risk than rotary dryers due to the shorter
retention time and lower operating temperature.

The final moisture is normally in the range of 10-20% wet basis. Retention time is in the order of
2-10s, providing the potential for rapid response to control signals. Size reduction is required
prior to the flash dryer. In flash dryers, oven dry throughputs ranging from 6.4-16 t/h and
evaporation is around 17 t/h. Flash dryer performance is summarized in Table 5- 1.
23


Sprout Dryer
Sprout dryers are similar to flash dryers, except they can handle slightly larger particles.
However, for a good cascading effect in the dryer, the particle size must be fairly uniform. Like
other direct air-heated dryers, heat recovery is difficult and expensive.

Moisture reduction is normally similar to that with rotary dryers, from 50-60% down to 20-40%
range, wet basis. Retention time is in the order of 2 minutes. Based on 4 cascade dryer
applications, Bruce and Sinclair (1996) summarized oven dry throughput rates ranging from
1.4-24.5 t/h. The dryer performance is summarized in Table 5- 1.

Belt Dryer
Belt dryers are better suited to take advantage of waste heat recovery opportunities because they
operate at lower temperatures than rotary dryers. Rotary dryers, for example, typically require
inlet temperatures of at 260C for drying, but more optimally operate around 400C. In contrast,
the inlet temperature of at least one commercially available vacuum dryer can be as low as 10C
above ambient, although more typically belt dryers operate at higher temperatures between about
90C and 200C. Because of their lower temperatures, belt dryers can even be used in
conjunction with a boiler stack economizer to take maximum advantage of heat recovery from
boiler flue gas. In this scenario, an economizer first recovers heat from the boiler flue gas, then
the exhaust from the economizer is used for fuel drying.

Their lower temperature also means that there is a lower fire risk. Emissions of volatile organic
compounds (VOCs) from the dryer will also be lower. An advantage of belt dryers over many
other dryer types is that the material is not agitated. This means there may be fewer particulates
in its emissions. On the other hand, fines may need to be screened out first and added back into
the dryer at a later point, since they can fall through the belts perforations.
5.2 Heat Recovery
Energy efficiency in air drying can be improved by using heat exchangers, recirculating exhaust
gases, multistage drying and heat pumps.

In heat exchangers, the heat is transferred from the exhausted gas through the wall of a heat
24

exchanger to the inlet gas. The exhaust gas is usually not saturated, so part of it can be
recirculated into the inlet of the dryer. Because the exhaust gas is generally warm, the energy is
not needed for heating recycled exhausted gas. Overall, the drying efficiency can be improved.
Besides, the exhaust gas, if high enough in oxygen, can be also used as preheated burner air.

The latent heat of water vapour in exhaust gases can be recycled to heat inlet material.
Wimmerstedt (1999) studied a case that a flue gas dryer was not co-located with other plants.
The dryer was operated with a dew point for exiting gas of about 80C. The gas can be used for
material pre-heating in a direct contact process. If there is a demand for the material to be
heating to a low temperature, a considerable part of the latent heat can be recovered. An
assessment of the heat consumption in one such plant gave a specific heat consumption of 3140
kJ/kg evaporated water. About 80% of the waste energy could be recovered for district heating
but the outlet water temperature was as low as 50C.

Multistage drying can be used when high inlet temperatures are a concern. Instead of diluting the
entire hot gas stream with cold air to reduce the temperature, some of hot gas can be introduced to
later stages of the dryer to boost the air temperature. As a result, less dilution air is used. A
run-around coil can be used where the physical layout of the dryer does not allow the exhaust gas
to be close to the inlet gas. A heat carrier, such as antifreeze, oil, or any kind of heating fluid is
first pumped through a heat exchanger coil in the exhaust gas duct, then through a heat exchanger
in the inlet air duct to release its heat to the inlet air. The disadvantage is that two heat
exchangers are needed, but this is sometimes cheaper than running extra duct work.

A heat pump is similar to a run-around coil. Because the heat pump uses a refrigerant and
compressor, it can recover part of the latent heat of vaporization by condensing or dehumidifying
the exhaust gas and then provide this heat to the inlet air at a higher temperature. Heat pumps
could improve the efficiency, but capital costs for the compressor can be very high with significant
compressor energy requirements. Heat pumps are also generally limited to providing heat at no
more than 60-66C.
25

5.3 Fire Safety in Dryers
Fire safety refers to precautions that are taken to prevent or reduce the likelihood of a fire. Fires
start when a flammable and/or a combustible material with an adequate supply of oxygen is heated
to an ignition temperature.
Combustible organic vapours are generally released from wood at temperatures from 200C,
where auto-ignition temperatures are generally 260-280C. Most dryers however can operate at
much higher temperatures to keep the air temperature greater than the biomass surface temperature.
This increases the drying rate, but also increases the fire risk in the dryer. For this reason, all
dryers are designed to minimize fire risk and are equipped with fire suppression systems.

In general, air drying has a potential high fire risk, because of the high amounts of oxygen in the
air supply. This can be reduced by limiting the amount of excess air or by recirculating exhaust
gases to the dryer inlet. Recirculation of exhaust gas also increases the thermal efficiency of the
dryer.

Flue gas dryers can operate at higher temperatures than air dryers, because flue gases have a lower
oxygen content (Intercontinental Engineering, Ltd, 1980). Compared with air or flue gas dryers,
superheated steam drying processes have a low fire risk. One risk however is the hot dried
biomass coming into contact with air after drying (Haapanen 1983, Wardrop Engineering, Inc.
1990).

As mentioned above, one important issue causing a fire hazard is the high temperature. One
effective method is a reduction in the operational temperature. For example, an operation
temperature of less than 100C could significantly reduce the likelihood of a fire. Various drying
processes at low temperatures will be discussed in Section 9. The maximum temperature that can
be used at the dryer inlet is limited by the burning temperature of dry wood in air around
260-290C. The flue gas can reach high temperatures, since they are depleted in oxygen relative
to air (typically 5-10% O
2
by volume).

5.4 Environmental Aspects
The exhaust gas from a biomass dryer may need to be treated before release to atmosphere. If
26

flue gas is used for drying wet fuels, the outlet gas from the dryer may contain SO
2
, CO
2
, CO,
particulate matter (PM) and hydrocarbons. Table 5-2 shows emissions data after secondary
cyclones for drying softwood.



Table 5-2. Softwood dryer emissions data (Wade, 1998).


A survey by Bruce and Sinclair (1996) indicated that for rotary dryers, total PM leaving the dryer
is lower than entering. In some cases, the reduction is as high as 50-80%. To achieve a
reduction in air emissions, the equipment is required to associate with the process. The first piece
of equipment after the rotary or flash dryers is a primary cyclone to separate the biomass from
exhaust gas stream. A set of multicyclones can follow the primary cyclone to remove some PM.
Cyclones, however, are not very effective for very small particles and thus a baghouse or wet
scrubber may be needed to remove small particles.

Some of particles may be removed in part by cyclones and wet scrubbers. However, condensable
volatiles can not be removed by dust collectors and appear in the stack emissions and result in
undesirable blue haze. At dryer temperatures higher than 180-220C, condensable resins and
organics will be released. After leaving the dryer, the condensed resins and organic acids form
aerosol, called as blue haze (Wade, 1998). These condensable organics are also counted as
particular matter (Bruce and Sinclair, 1996). The most effective method to control fine PM and
aerosols are wet electrostatic precipitators. The others include filter bed and electro-filter bed
scrubbers.

The volatile emissions mostly consist of monoterpenes. Monoterpenes are naturally emitted from
wood and have boiling points of 150-190C. The major components are -pinene and -pinene.
27

Photochemical reactions of monoterpenes with nitrogen oxides form low level ozone. Ozone is a
strong oxidant and a component of smog. In high concentrations, ozone is responsible for
impaired lung function in human populations, crop damage and is believed to be responsible for
forest damage in Europe and North America. Volatile emissions control can be achieved by
lowering peak temperatures, recycling exhaust gases leaving the dryer and reducing the amount of
fine material in the dryer feed material and its residence time.
5.5 Cost
In general, three kinds of cost information are available in the literature.

The first kind of cost information is the dryer equipment alone, i.e. flue gas dryer equipment:
rotary, cascade, and flash, and so on. The second kind of cost information is the complete costs,
including equipment and installation costs. The third one is the operating cost, such as costs by
electricity, water and gas, etc.

The costs may change according to the different dryer types, installation and the retrofit of the
dryer. During the course of this study, the costs for different dryers were identified:

Rotary Dryer
- Single pass rotary dryer = $20/kg/h, three pass rotary dryer = $18/kg/h.
- The dryer cost for 7.6 t/h wet wood chips = $323.000 (included installation cost);
- The dryer cost for 15-130 MW = $1.5-5.3 million (complete unit);
- Steam & Roger rotary single pass dryer = $26-47 /kg/h (complete unit);
- MEC rotary single pass dryer = $24-65 /kg/h (complete unit);
- Aeroglide rotary dryer = $17-31 /kg/h (complete unit);
- Heil rotary dryer = $32-88 /kg/h (complete unit);
- Biomass dryer = $38,000 /t/h(42/kg/h) (included installation cost);
- Biomass flue gas dryer for 55t/h = $5.4 million(complete unit);
The heat requirements were 3,000-8,1000 kJ/kg of water removed., with most estimated in
3500-4700 kJ/kg

Flash Dryer
- For a disk dryer, the cost was estimated at $5.4; the capital cost of a flash dryer was $18-35 /kg/h.
28

- The total cost of both the equipment and installation of a 15-130MW flash dryer for was
$550-1600 /kg/h.
- The cost of the dryer for bark = $350 /kg/h.
-There are two line Flakt flash dryer installed at the Assi Lvholmen Linerboard mill in Pitea,
Sweden. One unit supplies a lime kiln, the other, a recovery boiler converted to a power boiler.
The units cost 17.5 M$US in 1980, or the equivalent of 24.5M$US at the end 1995 for both the 6
and a 13 t/h capacity, or about 9.5 and 15MUS$ respectively. A second article describes a Flakt
flash dryer installation at E. B. Eddy Ltd, in Espanola, Ontario, which cost 14MCdn$ in 1986 and
handles about 18 t/h of wood waste (Wade, 1998).

Cascade Dryer
During the course of this study, technical articles providing capital costs for three cascade dryer
installations were identified:
Cascades Inc., East Angus, Quebec cost 36M$Cdn in 1992, or the equivalent of 32M$US at the
end 1995. Included were both cascade and flash drying and suspension firing of part of the fuel.
The cascade dryer with a throughput of about 9.0BDt/h accounted for about 5M$US;
Fletcher Challenge, Crofton, BC, cost 8.5M$Cdn in 1986, or the equivalent of 7.2M$US at the
end 1995. The cascade dryer had a throughput of about 36BDt/h.
Alabama River Pulp, Claiborne, AL, cost 6.3M$US in 1992, or the equivalent of 6.6M$US at the
end 1995. The cascade dryer had a throughput of about 32BDt/h.

Bruce and Sinclair(1997) summarized the costs for three kind of dryers and found that the total
installed costs for the complete dryer systems were very similar, if the wood fuel handling and
storage is excluded. This cost information is summarised in table 5-3. It should be emphasised
that the material handling equipment such as conveyors, feeders and bins is not included in any of
the cost information presented. These costs and retrofits where space is limited can add
substantially to the total cost of an installation.







29

Table 5-3. Capital Cost of Flue Gas Dryers
(a)
( Bruce and Sinclair, 1996).
Type
Moisture Content
In,% - Out,%
Equipment Cost
k$/t/h
Total Installed Cost
(b)

k$/t/h
Rotary 55 40 80 - 45 370 - 160
Sprouted 55 40 70 - 45 360 - 200
Flash 55 15 180 - 70 860 - 330
Notes: a - based on all the boiler flue gas entering at 300C and leaving the dryer at 105C.
b - the first value being for about 4 t/h, the second about 35 t/h.

Besides equipment and installation costs, the operating costs are important concerns. The main
components to the operating costs are those for power and maintenance. Power consumption
based on oven dry throughput are: 8-14 kWh/t for rotary dryers; 15-20 kWh/t for cascade; and
16-38 kWh/t for flash dryers, the latter depending on the size reduction required, which is a large
consumer of power. Size reduction power is highly variable depending on the equipment used,
size distribution of the feed, product size, type of material being pulverised, species and initial
moisture content. Summary data presented in a reference indicates a range of 10-100 kWh/t or
more.

In the absence of data encountered in the literature on the maintenance costs of equipment, an
allowance of 2% of total installed cost of the drying system equipment is suggested as the basis for
preliminary evaluation purposes.

5.6 Dryer Selection
The selection of dryers depends on a particular application. In general, water evaporation rate,
biomass property, biomass size, operation temperature and heating resource availability are
important in the selection. Meanwhile, environmental controls and safety are important
considerations in the dryer design.

For flue gas/air dryers, the selection of dryers is highly depended on the size of the biomass
material. For flash dryers, a small particle size is needed for moving air or steam to suspend the
particles. Triple-pass rotary dryers will accept larger material, but may experience plugging by
30

very large material. For large or variable material, a single-pass rotary dryer might be the best
option. In general, reducing the size of the material may be a good option for the drying process,
but it is an energy-intensive operation.

In the selection of dryers, the advantages and disadvantages of various dryers are always
considered. The significant advantages of rotary dryers include the fact that they are less
sensitive to material size, they operate at high temperatures to reduce drying time, their wide range
of evaporation rates and their easy installation. The major drawback is that these possess the
greatest fire hazard, since the high temperature operation is mostly applied in rotary dryers. Air
emissions need to be controlled. Heat recovery is difficult in rotary dryer. Flash dryers are
more compact and easier to control, but require a small particle size. Air emissions again need to
be highly controlled. Both flash and cascade dryers are used for high capacity water removal.
Belt dryers are currently adopted in low temperature operations, which present a lower fire risk,
reduced air emission and low energy consumption but require a large footprint. In general, rotary
dryers are the most commonly selected. Technical knowledge of equipment configurations,
installations and operations could provide useful information for new users. These are
summarised in Table 5-*.

Table 5-4. Summary of considerations in choosing a dryer.
Dryer
Type
Requires
small
particles
Heat
recovery
Fire
Hazard
Air
Emission
Drying
Temperature
(C)
Evaporation
(t/h)
Rotary No Difficult High medium 200-600 3-23
Belt No Easy Low Low 150-700 4.8-17
Flash Yes Difficult Medium High 160-280 2-41
Cascade No Difficult Medium medium 30-150 0.5-40

The capital costs of various dryers are often comparable. However, a belt dryer operated at low
temperatures may require less equipment for treatment of emissions; so for new installations the
overall cost may be less. The operation and maintenance costs of belt dryer are higher than for
other dryers. In general, multi-pass dryers are more complex than single-pass dryers and so have
greater operation and maintenance costs.
31

6. Superheated Steam Systems
Using steam to dry moist fuels has attracted recent interest, because of the advantages of the low
risk of fires, high energy efficiency and better environmental control.

The main components of a steam drying system are the fuel feeder, the flash dryer ducts or fluid
bed depending on the concept, cyclones, blower for recirculating the steam and a heat exchanger to
heat the superheated steam. The recirculating steam can be near atmospheric pressure or at a
higher pressure. The steam is in a superheated condition to provide the thermal driving force
necessary to evaporate the moisture in the fuel. The source of heat can be either very hot gases or
high pressure steam. Instead of air, superheated steam is used to suspend the solids and provide
the heat. Generally, the wet material is mixed with enough superheated steam to dry the material
and end with steam.

The superheated steam dryers have some key advantages compared to air dryers. No oxidation or
combustion reactions are possible. Steam dryers have higher drying rates than air and gas dryers.
Steam drying also avoids the danger of fire or explosions and allows toxic or valuable liquids to be
separated in condensers. However, the systems are more complex and even a small steam
leakage is devastating to the energy efficiency of the steam dryer.

6.1 Dryer Type
In the 1980s, an energy efficient steam drying technique using superheated steam or turbine
backpressure steam was developed in Sweden by Modo-Chemetics for drying wood pulp, bark,
agricultural products, peat and other biomass materials under pressure. Since that time, different
forms of steam dryers have been developed, tested and in some cases, commercialised. This
study introduces different approaches to steam drying. Steam drying technologies may be
atmospheric or pressurised, have short or long residence times and be large or small. Simplified
block diagrams of various types are presented in Fig 6-1.
32


a. Basic IVO steam dryer b. IVO steam dryer with fluidized bed


c. Niro steam dryer d. Luri steam dryer
Figure 6-1. Simplified block diagrams of various types of superheated steam dryers.

The first is the basic IVO dryer made by Imatran Voima Oy (IVO), where biomass material is
mixed with recycled superheated steam, as shown in Figure 6-1 a. The superheated steam and
biomass pass through a flash tube and the solids are separated from the steam in a cyclone, the
same as in a flash dryer. Most of steam is recycled through a fan to provide the force to suspend
the solid material and then the steam passes through a heat exchanger to increase its temperature.
The extra steam can be condensed to recapture the latent heat, compressed to a higher temperature,
or with high pressure operations, steam can be injected into a gas turbine to increase the power
output (Hulkkon et al. 1991).

The second IVO superheated steam dryer is a bed mixing dryer, as shown in Figure 6-1 b. It can
be used with a fluidized bed gasifier or boiler. Some of the hot bed material from the combustion
chamber is mixed with the wet biomass in steam. The sensible heat from the bed material
evaporates the water from the fuel. The fuel can then be fed, along with the bed material back
into the process, while the steam can be recycled, with the excess steam being used for other
applications (Hulkkonen et al., 1991).

Niro of Denmark has commercialised its superheated steam fluid bed dryers and have numerous
33

installations. The fluid bed operates under pressure of 200-300kPag (29-44 psig) by circulating
steam in a closed vessel. Wet material is fed into the first of 16 vertical cells arranged around a
centrally located heat exchanger. In the cell, the material is kept in suspension by superheated
steam entering from the bottom. The material moves successively through each of the 16 cells and
is finally discharged by a screw conveyor. Niro installations are mainly applied to the drying of
sludge and other by product materials in the agricultural industry as a part of reprocessing.

The Lurgi technology from Germany uses a fluidised bed consisting itself of the material being
dried. The material to be dried must be capable of assuming and maintaining a granular form, but
many materials, according to Lurgi, exhibit this property, including: peat, lignite, sewage sludge,
pulp and paper mill sludge, agricultural by products, ore and mineral sediments. Heating of the
wet material and reheating the superheated steam is done using a heat exchanger tube bundle
which is immersed in the bed material and provides very high rates of heat transfer.

A steam drying system developed by Modo-Chemetics for wood waste fuels was adapted from
earlier work on pulp drying and was promoted in the early 1980s for fuel drying. Only one such
system was installed, and is no longer in service. This technology, similar to Stork's, is no longer
being offered.
6.2 Advantages and Disadvantages
The main advantage of superheated steam dryers is that the latent heat of vaporization from drying
can be recovered and no heat losses occur compared with heating air for drying. There are
normally no air emissions from superheated dryers because all the vapours, including organic
species, are condensed. Nevertheless, wastewater treatment should be considered. Superheated
steam dryers have higher heat transfer rates and faster drying. The presence of an inert steam
atmosphere has no fire hazard. The IVO mixed bed steam dryer does not need heat exchangers
for drying. A high-pressure IVO dryer integrated with a gas turbine eliminates the wastewater
stream by combusting the organics in the turbine.

The heat source and temperature for drying are important for dryer selection. Compared with
flue gas, a process stream for heating exchange is energy efficient, but it requires the capital
investment in heat exchangers and the interactions between the dryer and process. Superheated
steam dryers require a high-temperature heat source. The selection and design should consider
34

what excess heat is available in the system, and then design a suitable drying system. If not, a
burner can be installed with an auxiliary fuel source to provide the heat for drying.

The disadvantages include the small particle size required for mixing steam and particles, high
capital costs for a stainless steel pressure vessel and wastewater treatment.

In steam flash dryers, the shortest distance of the fuel particles must not exceed 4-5mm
(Wimmerstedt, 1999). Particle lengths up to 30mm are acceptable. The steam drying system is
therefore needed to warm the wet fuel before the dryer. The power for the fan and the screws is
needed for transportation of the fuel in and out from the dryer at elevated pressures. The other
reported problems include corrosion and clogging, as well as leakage through the exiting plug
screw.

The fluid bed dryer is claimed to accept particles up to 20 mm. In the dryer installed at the pulp
mill, the fuel had a high content of bark strips and shaves, which passed the strainer causing severe
plugging of the dryer. To solve the problem an extra crusher had to be installed before the dryer.
The fan power is about 25% lower in the fluid bed dryer compared to the flash dryer.

Sensitive parts of the fluidized dryers have to be clad in stainless steel. The most severe
problems have been with the cell feeders. They have had to be renewed several times, and have
limited the capacity to the design value.

In a steam drying plant the material temperature corresponds to the saturation temperature at the
prevailing pressure which is 130-150C. The steam is condensed after the dryer. The inert
gases from the condenser could be incinerated. The condensate contains terpenes, which can be
decanted from the water phase and recovered. The condensate also contains COD and phenolic
compounds. In order to remove these compounds, the condensate at the pulp mill was discharged
to the spill water tanks, evaporated and burnt together with the black liquor. This is of course a
safe method, but on the other hand it results in a waste of energy.

Though steam drying is not limited by concerns of fires and flue gases causing corrosion, there are
process constraints on such systems. The use of excessively high temperatures in steam dryers
will result in the volatilisation of part of the fuel, with a high contaminant level in the resulting
condensate.
35

6.3 Capacities of Dryers
Besides the advantages and disadvantages of each dryer, the capacity and costs in industrial
applications are critical parameters to determine the selection of dryers. Table 6-2 lists the
capacity and operation parameters of industrial dryers (EPRI).

Table 6-2. Capacities and performances of steam dryers (Bruce and Sinclair, 1996).
Company Dryer
Evaporative
Capacity,
t/h
Operating
Pressure in
Dryer, kPag
Steam Pressure
of Heat Source,
kPag
Average
Residence
Time, s
IVO Flush 0.7(1) 300-2200 - 2-3
IVO Flush 5.6(2) ~0 - 2-3
Niro Fluidized bed 1-40 200-300 1500-2800 300
Lurgi Fluidized bed >150 15 - 25 400 - 900 -
1 - One 0.7 t/h pilot plant, and range of pressures tested
2 - One demonstration plant of 5.6 t/h

6.4 Performances
The superheated nature of the recirculating drying steam is achieved by a heat exchange with the
heat source. From Lurgi's literature, the high pressure steam required for lignite and peat steam
dryers is 400-900kPag, giving dryer steam temperatures of 120-150C, assuming a 40C
difference. For agricultural and wood residues, the corresponding pressure is 1000-2500kPag,
giving dryer steam temperatures of 144-186C, again assuming a 40C difference.

The minimum temperature of the recirculating drying steam is a function of the operating pressure
of the dryer. For efficient use of the dryer capacity, the steam must remain superheated. A 20C
above saturation temperature is commonly accepted, though equipment suppliers will have their
own criteria.

The throughput rate, evaporation capacity and other design and performance data of a Niro system
described in the literature are presented in Table 6-3.

36

Table 6.3. Design and performance data of a Niro A/S steam dryer.
Evaporation (t/h) 27.2
Capacity (t/h) 10.8
moisture, Feed % 70.8
moisture, Discharge % 10
Thermal requirements:
without heat recovery (GJ/t_evap) 3.45
with latent heat recovery (GJ/t_evap) 0.4
Blower power (kWh/t_evap) 67


6.5 Heat Recovery
In superheated steam drying, the latent heat of vaporization is easier to recover because the water
vapour that leaves the fuel is not diluted by air, so it can be condensed directly to recover the heat.

Depending on dryer and plant configuration, there are several possibilities for heat recovery. If
power plants could provide hot water for heating, superheated steam processes can be operated at
atmospheric pressure. If process steam is required, the dryer must either be operated at a higher
pressure or the steam from the dryer must be compressed to increase the temperature. The steam
from the dryer can be used directly or it can be condensed on the outside of the boiler tubes to
produce clean steam without any impurities.

In the case of a steam dryer incorporated in a process industry or a district heating plant, the steam
from the dryer can be recovered at high temperatures and can be further used in a process industry
or in a DHS at high temperatures, as shown in Figure 6-2 (Wimmerstedt, 1999). From a
thermodynamic point of view, the heated steam to the dryer should be extracted from the turbine at
as low a pressure as possible, to ensure maximum cogeneration. The waste heat from the dryer is
mainly used for feed water heating, which gives more cogenerated electricity than using it for
district heating (Wimmerstedt, 1996). The IVO Bed Mixing Dryer could be included in this case.
The heat to the dryer is supplied directly from the boiler via the hot bed material. The product
steam is used for heating the district heating water. The entire recovered heat load reduces the
basis for cogeneration. The coupling accordingly gets the same function as increasing the
37

coefficient of efficiency for the boiler.


Fig. 6-2. Steam dryer with heat recovery.

6.6 Cost
Capital Costs
According to the report of Bruce and Sinclair (1996), IVO High Pressure Steam Dryer system total
installed costs are expected to be of the order of 300,000-400,000 $/t
evap
/h (price in 1996), which
includes a pulveriser for some size reduction.

According to the report of Wade (1998), the total capital cost of MoDo type steam dryer was $ 4.5
million; the cost of IVO system with heat exchanger was $7 million; for a disk dryer, the cost was
estimated at $5.4. The capital cost of flash dryer was $18-35 /kg/h. The total cost of both
equipment and installation of flash dryer for 15-130MW was $550-1600 /kg/h. The cost of the
dryer for bark was $350 /kg/h.

Operating Costs
The main components to the operating costs are cost of power and maintenance. Power
consumption data is very limited, but literature from Niro and IVO suggest a range of 20-45 kWh/t
38

throughput excluding the pulveriser; the latter could add another 10-50 kWh/t.
6.7 Environmental Aspects
In steam drying, contaminated condensates are formed when the water and other less volatile
vapours evaporated from the wet fuel are condensed after drying and the energy is recovered.

The organic load or strength of the contaminants is found to be surprisingly high and increases
rapidly with increasing drying temperature. Bruce and Sinclair (1996) reported a tripling of BOD,
COD and TOC loads from peat as temperatures were increased from 110C to 130-140C, while
the same parameters doubled for forest residues as temperatures were increased from 140C to
160C. For peat, the dry weight loss increased from 1% at 190C to 10% at 350C. Waste water
data specific to steam dryers is presented in Table 6-4.

Mackie (1995) reported on the mild distillation of peat at 180-220C for 30 minutes showed
12-14% of the dry weight of peat as dissolved solids in the condensates which had a COD of
14,000 mg/L. Hulkkonen (1994) found that for steam drying operating at 2300kPa / 220C, the
TOC was found to be 1100-1500 mg/L for crushed wood chips and 450-900 mg/L for sawdust.
Lipophilic compounds, which are potentially tacky at drying temperatures, were measured at
2300-4400 mg/L or 0.20-0.45% of the dry weight of which 70-90% were fatty and resin acids.
The condensates from the pressurized drying experiments were found to contain lower amounts of
lipophilic extractives than those from atmospheric drying experiments.

These condensates require treatment before release to water courses. Conventionally, the
treatment would comprise neutralization, decantation or filtration of suspended solids and
biological treatment for reduction of oxygen demand and toxicity. Alternatively, steam stripping
may be found efficient to concentrate the impurities. This would allow recovery of the energy in
the organics rich stripper off gas stream for firing in the boiler.






39


Table 6-4. Wastewater characteristics from steam dryers.


The non condensable components in the dryer steam include rich-terpene, CO
2
and lesser amounts
of H
2
, CO, CH
4
, and C
2
- C
4
(Fagerns, 1996). Those components can be disposed of by firing in
the boiler or furnace receiving the dried fuel (EPRI).

7. Recent Development in Low Temperature Heat Sources for
Biomass Drying
In thermal drying, moisture in wet biomass is evaporated by introduction of a hot gas (air)/steam
stream. With typical temperatures averaging 350C, the dried product contains 70-90% solids by
weight. Typical thermal drying installations include unit operations like blending of wet and
dried solids, drying, dried solids handling and exhaust gas cleaning. Most designs require wet
scrubbers and dry solids collectors to prevent particulates and odours from polluting the air. As a
result, most thermal drying plants are large, complex systems associated with high capital
investment and operating costs. Furthermore, control and handling of dust emissions and
high-fire risks are a problem with all thermal dryers.

Biomass drying using low temperature heat sources offset the high capital and operating costs
associated with traditional drying methods. Under a low-temperature operation environment,
dust emissions and fires risks could be significantly reduced. As reported so far, industrial dryers
40

available for low temperature heat sources are provided by Swiss Combi, Bruks Klckner,
Mabarex (Thermal Energy), Andritz Fiber and Svensk Rkgasenergi (SRE), among others.
Those dryers have capacities of 500-4000 kg/h under operation temperatures of 30-110C. The
following section introduces some typical applications of dryers using low temperature heat
sources. Besides, microwave dryers are also briefly introduced as a potential commercial dyers.

7.1 DRY-REX
Thermal Energy International Ltd has installed a low temperature biomass drying system,
DRY-REX, at Uniforet pulp and paper mill in Quebec to dry sludge into high efficiency biofuel.
The system processed 200 wet tons of sludge a day containing 22% solids (thus 78% moisture),
evaporating 5 tons/hr of liquid from the sludge at low temperature with virtually no emissions.

The DRY-REX dryer is an ambient temperature system, which uses a patented, two-step,
integrated granulating-drying process (Barre and Bilodeau, 1999). The feed stream from the
dewatering system is first processed in a granulator on top of the dryer. Granulation reduces wet
residue bulk, optimizes particle size and solids distribution for easier handling and exposes
maximum surface area for fast drying. The dryer consists of a totally enclosed structural frame
with several stacked drying zones, each with its own solids conveying system and air distribution
assembly, as shown in Fig. 7-1. The continuous tunnel-type convection dryer uses a vacuum,
forced-air stream and air temperature above 5C as the main driving force. The 50% of the drying
air is recirculated to reduce the quantity of exhaust air. The residence time is 15-30 minutes for
sludge and wood chips. Thermal energy for water evaporation is 50-150 kWh/t water evaporated,
much less than other dryers, as shown in Table 7-1. The capital costs, which include supply,
installation, operation, maintenances and energy, are 2.4-6.5 U.S.$/wet ton (in 1999) for sludge
and 0.5-1.5 U.S.$/wet ton (in 1999) for bark. The identified problem was that temperatures
above 5C limited the absorption of water. Barre and Bilodeau (1999) suggested that the
temperature should be raised to 20C.

41


Figure 7-1. A sky view of the compact dryer.

Table 7-1. A comparison of operating temperatures and thermal energy for various types of dryers.


Drying tests in pilot dryers have been done in 25 pulp and paper mills on most types of residue
using a trailer-mounted demonstration unit. Table 7-2 summarizes the typical performance of the
DRY-REX in the most common environmental pulp and paper applications. In pulp and paper
sludge applications, the significant benefits are the savings in operating costs as a result of the
lower consumption of fossil fuel. There is also a net gain for the environment because it reduces
greenhouse gas emissions caused by burning fossil fuels. In general, drying sludge to high solid
content leads to a reduction of up to 60% in volume and weight. As a granulator was used in
DRY-REX, some of the dewatering processes (belt press and vacuum filter) can be eliminated
from the existed plant. In some plants, sludge dewatering consists of a belt press followed by a
screw presses in series. After pressing, the product was extremely dusty and contained 50%
water. The pilot dryer was fed directly from the belt press at 30 wt.% solid and delivered 80 wt%
42

solid product, enabling mills to eliminate the dust problem and to bypass the screw presses.

Table 7-2 Typical results in the pulp and paper industry.


For drying bark, tests were done at several mills. The pilot dryer in the tests was fed at an
average rate of 42 wt% solids (as low as 27.6wt% solid) and delivered on average 75 wt% solids.
There were no significant differences between drying raw and shredded bark, but bark particles
should be limited to 75x75x20 mm or less to prevent potential accumulation and obstructions in
the dryer. Drying tests were also done on mixtures of bark and sludge with various mixing ratios.
The results revealed that drying bark alone and drying the mixture of sludge and bark were
essentially identical. Table 7-3 illustrated the effects of feeding dried bark into the existing power
plant.








43


Table 7-3 Drying converts wet residuals into a profitable energy that reduces consumption of fossil
fuels and lowers the emission of greenhouse gases.


7.2 SRE- Renergi LTD Dryer
Svensk Rkgasenergi (SRE), a subsidiary of Opcon AB, the energy and environmental technology
Group, signed an agreement concerning delivery of its new energy-efficient low-temperature drier
for biomass System Renergi LTD, to Corbat SA, a Swiss sawmill. SREs low-temperature drier
will be used to produce pellets.

The Renergi LTD low-temperature dryer can use heat at temperatures of 50C to dry sawdust,
chips, bark and similar material. Because of the low temperatures the heat can be recovered from
waste energy sources. This also saves significant the operating costs. The Renergi LTD dryer is
based on the counter-flow principle. Material is fed from the top of the dryer and removed at the
bottom. The heart of the dryer system consists of two drying cells. The task of the drying cell is
to provide optimal exposure to the sawdust so that its water content can be effectively taken up by
the dry air that passes over the it. The wheels on the dryer cell consist of 16 shovels that are
designed to spread the sawdust into the air. This ensures efficient exposure of the solids to the air,
which can absorb the water at a high rate. The long-term retention time in each drying cell
guarantee homogeneous drying.

The fan sucks in the outside air via the air battery LB1, as shown on Figure 7-2, which raises the
air temperature. The dry and warm air offers good conditions for effective drying. The water
content of the sawdust is absorbed in TC1 by the dry air until the air is saturated with water and
44

leaves the drying cell with a lower temperature. Air passes through an air filter (LF1), which
removes sawdust before the air is heated once again via LB2. Now the air can perform a
secondary drying process in TC2, leaving the dryer via the LF2 air filter. Once this drying batch
is ready, the process returns to emptying and refilling of dryer cells. Table 7-4 shows the
performance of this dryer. There is no further information regarding its operation and
maintenance.


Figure 7-2. Functional diagram and cell structure of Renergi LTD Dryer.

Table 7-4. Performance of the RENERGI LTD low temperature dryer (Incoming fresh air to dryer:
Temperature: 0C, Water content in per kg of dry gas: 2 g/kg, Enthalpy: 5 kJ per kg of dry gas /
Saw dust: 55 % moisture content in incoming saw dust, 10 % in outgoing saw dust.


45

7.3 Microwave
Conventional drying of wood is the most energy-intensive and costly process. Conventional
wood dryers function under the basis of convective heat transfer from hot air to the surface of
wood followed by conductive heat transfer from the surface to the centre of wood. These dryers
require considerable amounts of energy and long drying times in order to evaporate water from the
wood. Unlike the conventional dryers, where heat is applied externally to the surface of the
material, microwave simultaneously heats the bulk of the material. When properly designed,
microwave drying systems have several advantages over conventional methods including a
reduction in the drying time, high energy efficiency, improvements in product quality for various
industrial applications, a reduction in fire hazards and lower air emission. Microwave drying of
wood products, however, has not been used to a larger extent in wood industries mainly due to the
insufficient knowledge of the complex interaction between wood and process parameters during
drying as well as the higher investment expenses.

The technology of microwave heating and drying in the field of wood products started to be used
in the early 1960s. Microwave drying and heating processes can be operated at temperatures
below 100C, which is beneficial in terms of energy savings, safety operations and emission
control. In addition, drying wood with the use of microwaves is faster than conventional drying.
Antti (1999) has shown that the microwave drying of pine and spruce is 20-30 times faster than
with conventional methods.

The wood drying rate through using microwaves mainly depends on the frequency of microwave
irritation, the wood material, the temperature and the wood sizes (Hansson, 2007). At a
frequency of 2.4 GHz, a drying rate of 0.4 fractional moisture per hour could be reached for boards
of spruce and beech (Hansson, 2007) and a 25-mm-thick Korean red pine could be dried in less
than one and half hours (Lee, 2003). With a frequency of 915 MHz and a manipulated
microwave input power and hot air, a 25-mm-thick pine plank could be dried in less than three
hours (McAlister and Resch, 1971). A prototype continuous microwave dryer for softwood
structural lumber could dry 50-mm-thick hemlock and Douglas fir in 5-10 hours (Hansson, 2007).

It is known that drying by microwave technology is a high energy consumption process even
though it saves thermal energy. To improve energy efficiency, Turner (1994) pointed out the
suitability of combined microwave and convective drying. Lee (2003) investigated the
46

performance of combined microwave and convective drying of wood. In his study, Korean red
pine was selected as material. To study efficiency, Lee (2003) compared specific moisture
evaporation rates (SMER). In general, the average value of SMER in microwave drying was
found to be about 1 kg/kWh (Lee, 2003). As shown in Table 7-5, evaporation rates were very
low if combined microwave and convective heat was used. To improve energy efficiency, the
exhaust gas should be recycled to inlet.

Table 7-5. Performance characteristics of combined microwave and convective drying Korean red
pine at a constant hot air energy input of 1kWh per one hour without heat recovery.


For wood drying using microwaves, the effect on wood hardness and structure is a primary
concern. Vongpradubchai and Rattanadecho (2009) investigated wood properties after
microwave treatment. SEM results demonstrated that microwave dried specimens has a better
micro structure arrangement because of uniform energy absorption, heat and moisture distribution.
In addition, the microwave heating offers better mechanical properties with high strength and little
deterioration in its long term performance with higher quality than conventional methods.

8. Conclusions
The use of dried biomass could provide significant benefits to operation of boilers. It will increase
boiler efficiency, lower emissions and improve operation. The drying mediums used are air, flue
gases and superheated steam. The dryers are commonly rotary dryers, flash dryers, belt dryers
47

and fluidized bed dryers. Drying by air and flue gases mostly takes place in rotary, flash, belt and
cascade dryers. Superheated steam is mostly used for drying biomass in flash and fluidized bed
dryers.

The selection of a dryer depends on its particular application. In general, water evaporation rate,
biomass properties (including size), operating temperature and heating resource availability are
important in the selection. Meanwhile, environmental controls and safety are important concerns
in the dryer design.

Rotary dryers are the types used with air or flue gases. The significant advantages include less
sensitivity to material size, a wide range of evaporation rates, easier installation and an abundance
of applications. The high temperature operation is mostly applied in rotary dryers to accelerate
drying. The significant drawback is the fire risk and significant air emissions due to the high
temperature operation. Flash dryers are more compact and easier to control compared to rotary
dryers, but require a small particle size. Both flash and cascade dryers can be used in high
capacity water removal. Heat recovery is difficult in rotary dryers, flash tube dryers and cascade
dryers. Belt dryers are currently adopted in low temperature operations, as they pose less of a fire
risk, emit fewer air pollutants and have a low energy consumption, despite their requirements of a
large footprint.

Compared with air dryers, the main advantage of superheated steam dryers is that the latent heat of
vaporization from drying can be recovered. There are normally no air emissions from
superheated dryers because all the vapour, including organic species, is condensed. The presence
of an inert steam atmosphere has no fire hazard. The disadvantages include the small particle
size that is required for mixing, the high capital costs for a stainless steel pressure vessel and the
wastewater treatment. Superheated steam dryers require a high-temperature heat source. Most
superheated steam dryers were constructed near process and power plants, since hot water and flue
gases provide the resources of high temperature and energy. As a superheated steam dryer, the
IVO mixed bed steam dryer does not need heat exchangers for drying. A high-pressure IVO
dryer integrated with a gas turbine eliminates the wastewater stream by combusting the organics in
the turbine.

Drying biomass at low temperatures (<100C) is a promising technology, mainly due to the
reduced environmental pollution and fire hazards. The commercial dryers available for biomass
48

drying are belt dryers and two cell dryers. Both are operated under vacuum conditions.
Microwave dryers may need more development and research before more widespread applications
for biomass drying.

9. References
Antti, A. L. (1999). Heating and drying wood using microwave power, Ph.D. Thesis. Vol. 35.
Lule University of Technology, Division of Wood Physics.

Barre, L. and M. Bilodeau(1999), Drying residuals at low temperature with the Dry-Rex dryer,
Pulp and Paper Canada, 1999

Berghel, J., Lars Nilsson, Roger Renstrom (2008), Particle mixing and residence time when
drying sawdust in a continuous spouted bed, Chemical Engineering and Processing 47: 12461251

Bruce, D. M. and M. S. Sinclair (1996). H.A. SIMONS LTD, Thermal Drying of Wet Fuels:
Opportunities and Technology, Final Report(TR-107109 4269-01)

ELECTROWATT-EKONO (UK) LTD 2003, Maximising the Potential of Wood use for Energy
Generation in Ireland.

FBT, Inc., (1994), Fluidized bed Combustion and Gasification: A Guide for Biomass waste
Generators, Southerneastern Regional Biomass Energy System. Work performed by FBT, Inc.,
Chattanooga, TN

Fredrikson, R.W. (August 1984), Utilisation of Wood Waste as Fuel for Rotary and Flash Tube
Wood Dryer Operation. Biomass Fuel Drying conference Proceedings; Aug. 8, 1984, Superior,
Wisconsin. St. Paul, MN: Office of Special Programs, University of Minnesota; pp.1-16

Hansson, Lars (2007): Microwave Treatment of Wood. Doctoral thesis, Lule University of
Technology, Skellefte, Sweden.

Haapanen, A.P.; Heikkila, L.; Ljias, M.; Valkamo, P.; (1983), Enso uses Flash-directed, pulverized
49

back to replace coal as boiler fuel Pulp & paper; 77. 70-71

Holmberg, Henrik (2007): BIOFUEL DRYING AS A CONCEPT TO IMPROVE THE ENERGY
EFFICIENCY OF AN INDUSTRIAL CHP PLANT. Helsinki University of Technology

Huhtinen M, Hotta A. Combustion of Bark. In: Gullichsen J, Fogelholm C-J, editors. Chemical
Pulping (1999). Helsinki, Finland: Papermaking Science and Technology; p. 203-301.

Hulkkonen, S.; Raiko, M.; Aijala, M. (1991), High Efficiency Power Plant Processes for moist
fuels IGTI. Vol.6 avaiable from ASME Book No. I00313-1991

Intercontinential Engineering, Ltd. (1980), Study of Hog Fuel Drying Systems. Canadian
Electrical Association. Prepared by Intercontinential Engineering, Ltd.

Lee , H. W. (2003). Combined Microwave and Convective Drying of Korean Wood Species, 8th
International IUFRO Wood Drying Conference

Mackie, K. L. (1995) The nature and control of solid, liquid and gaseous emission from
thermochemical processing of biomass, New Zealand Forest Research Institute, Rotorua, New
Zealand. Report No. ES 94/2 of the Task X, Activity & - Environmental Systems, International
Energy Agency

McAlister, W. R., and Resch, H. (1971). Drying 1-inch ponderosa pine lumber with a combination
of microwave power and hot air. Forest Prod. J. 21(3):2634.

Shao, Yuanyuan, Jesse Zhu, Fernando Preto, Guy Tourigny, Jinsheng Wang, Chadi Badour,
Hanning Li, Chunbao (Charles) Xu (2009). CHARACTERIZATIONS OF DEPOSITED ASH
DURING CO-FIRING OF WHITE PINE AND LIGNITE IN FLUIDIZED BED COMBUSTOR,
Proceedings of the 20th International Conference on Fluidized Bed Combustion, 1041-1047

Spets JP, Ahtila P.(2004) Reduction of organic emissions by using a multistage drying system for
woodbased biomass. Drying Technology: 22: 541-561.

Stahel, Roger, Marco Baumgartner, LOW-TEMPERATURE DRYING, 17th BIOMASS
50

CONFERENCE HAMBURG

Turner I. W. (1994). A study of the power density distribution generated during the combined
microwave and convective drying of softwood. 9
th
International Drying Symposium, Gold Coast,
Australia, August 1-4.

Vongpradubchai S. and P. Rattanadecho (2009). The microwave processing of wood using a
continuous microwave belt drier, Chemical Engineering and Processing 48: 9971003
Wade A. Amos, Report on Biomass Drying Technology, NREL 1998

Wimmerstedt R. (1995) Drying of peat and biofuels. In: Mujumdar AS, editor. Handbook of
industrial drying. New York: Marcel-Decker; p. 809-859

Wimmerstedt R., J. Hager (1996), Steam drying: modeling and applications, Drying Technol. 14 (5)
10991119.
Wimmerstedt R. (1999) Recent advances in biofuel drying, Chem. Eng. Process. 38 (46), p
441447.

You might also like