Você está na página 1de 20

Chapter 21

Dryhouse technologies and DDGS production Reprinted from The Alcohol Textbook 303 5th Edition 2009 with permission from Lallemand Ethanol Technology and Nottingham University Press

Dryhouse technologies and DDGS production

D.A. MONCEAUX1 AND D. KUEHNER2


1 2

AdvanceBio, LLC, 5405 Dupont Circle, Milford OH 45150 USA (monceaux@advancebiollc.com) Barr-Rosin, Maidenhead, Berkshire, UK (dominique.kuehner@geagroup.com)

INTRODUCTION
The cereal grains used for starch-based ethanol production such as corn, wheat, barley, rye and grain sorghum (milo) typically contain from 50 to 75% w/w starch (dry solids) (Table 1). Between 25 and 50% w/w of the grain remain after fuel ethanol has been produced. The whole stillage, containing both dissolved and insoluble nonfermentables and nondistillable microbial by-products, is rich in nutrients, fibre, protein, lipids and yeast and has traditionally been incorporated into animal feed rations.
Table 1. Composition of cereal grains used in alcohol production. Component Starch Protein Fat Ash Corn 6572 912 4.5 1.0 Content (weight, % of dry matter) Wheat Barley Rye 6770 1214 3.0 2.0 5264 1011 2.53 2.3 5565 1015 23 2.0 Sorghum 7275 1112 3.6 1.7

removed by centrifugation), distillers dried grains (DDG), distillers dried grains with solubles (DDGS), thin stillage (backset) and/or syrup (concentrated thin stillage, which may be sold separately). Excepting wet grains and thin stillage that must be used near the site of production, these products can be costeffectively stored and shipped to distant animal feed ration markets.

DRY-GRIND ETHANOL PROCESS OVERVIEW


Historically, ethanol production from starch-bearing cereal grains used one of two general categories of process technologies: wet milling or dry grind. Recently, dry-grain fractionation processes, similar to technologies that were in practise before the advent of corn wet milling, have been re-introduced to the dry-grind fuel ethanol industry. Dry-grain fractionation processes enable the ethanol producer to extract additional value from the non-ethanol coproduct by separating it into a broader spectrum of products for both conventional and developing markets. Currently, the majority of fuel ethanol production facilities not using wet milling employ conventional dry-grind technologies, with variations of the following major unit operations:

The dryhouse is where these valuable animal feed components are recovered from the ethanol production process. The sole purpose of dryhouse operations is to remove water from whole stillage in an efficient and mechanically reliable manner. This process yields a number of possible concentrated products distillers wet grains (only some water is

304 D.A. Monceaux and D. Kuehner Milling Mashing Cooking Liquefaction Saccharification Fermentation Ethanol process unit operations are often managed as separate front-end and back-end processes, a practise derived from the industrys early association with the distilled spirits industry. Numerous process and energyintegration strategies continue to blur the distinction, as technology changes drive greater process efficiencies and capital cost reductions.

Front-end processing

Distillation Alcohol dehydration Centrifugation* Evaporation* Drying*


* Dryhouse unit operations

Back-end processing

DRYHOUSE PROCESSES OVERVIEW


Two basic processes (solids separation and dehydration) and three unit operations (centrifugation, evaporation and drying) make up the dryhouse in a modern, largescale dry-grind fuel ethanol plant. Other unit operations have been used on a limited basis. The objective is to concentrate the solids fraction of whole stillage using a combination of mechanical separation and thermal processes. A wide variety of technologies are available. Those selected must deliver mechanical reliability and thermal efficiency at a moderate capital investment. The resultant product is 90% dry matter DDGS, made by maintaining properties of the original, approximately 15% dry matter, whole-stillage feedstock. Table 2 provides information on DDGS composition. In this discussion, DDGS is assumed to be the final product made in the dry mill.

These eleven unit operations are typically grouped into the two general process areas, as shown in Figure 1. The front-end processes convert the starch-rich cereal grain into ethanol as efficiently as possible. The ethanol-rich beer that is produced in the front-end operations is recovered in the back-end distillation process and dehydrated to produce fuel ethanol. The residual suspension, called whole stillage, rich in nonfermentable dissolved and suspended solids, is processed in back-end dryhouse operations, usually producing distillers dried grains with solubles (DDGS) and thin stillage/backset.

Figure 1. Dry-grind block flow diagram.

Dryhouse technologies and DDGS production 305


Table 2. A typical analysis of corn DDGS. Component Dry matter Crude protein Fat Acid hydrolysed fat Fibre Ash Nitrogen-free extract Carbohydrates Acid detergent fibre Total digestible nutrients Calcium Phosphorus Potassium Magnesium Sulphur Sodium Chloride Zinc Manganese Copper Iron Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Phenylalanine Theonine Tryptophan Valine Anonymous (2008). Concentration 89.18% 30.03% 10.86% 11.06% 7.22% 5.97% 44.73% 51.96% 13.72% 86.49% 0.07% 0.74% 1.01% 0.3% 0.67% 0.18% 0.19% 61.63 ppm 18.40 ppm 6.29 ppm 126.00 ppm 1.32% 0.82% 1.17% 3.55% 0.91% 0.64% 0.66% 1.51% 1.11% 0.24% 1.57%

the fuel ethanol industry has experienced only two periods of significant growth between protracted periods of limited new plant construction (this has restricted technology development), alternative large-scale dryhouse processes have not been demonstrated on a commercial scale and project finance strategies have suppressed the implementation of new, innovative technologies. Most improvements in dryhouse operations and energy efficiency have been the result of incorporating interprocess unit operations, energy-integration strategies and technology breakthroughs into front-end processes. Foremost has been the dramatic reduction in stillage volumes that has resulted from increasing beer ethanol content. Ongoing improvements in fermentation technology and yeast metabolism have brought about dramatic increases in per cent v/v ethanol content in fermentors. In the early 1980s, corn dry-grind fuel ethanol plants considered 8% v/v ethanol beer acceptable. Today, a number of plants claim consistent operation at ethanol concentrations above 20% v/v (Figure 2). The impact of this very high gravity fermentation is a significant reduction in the related decanter centrifuge and evaporator duty and in plant investment.

Dryhouse mass balance


The increased demand for renewable fuel has led to a corresponding increase in dry-grind fuel ethanol plant production capacity. While new plant annual production capacities averaged less than 30 million denatured gallons ten years ago, today most new dry-grind plant

Nutritional values expressed on 100% dry matter basis.

Dryhouse technologies have changed little over nearly half a century because
100 Relative stillage flow rate (%) 80 60 40 20 0

10

12

14

16

18

20

22

Volumetric beer ethanol content (%)

Figure 2. Relationship between beer ethanol content and stillage flow rate.

306 D.A. Monceaux and D. Kuehner capacities range from 60 to 100 million US gallons per year, and above. To successfully manage dryhouse operations, it is important to know the mass balance (or material balance) the quantities of materials that need to be processed. Process simulation modelling of a corn drygrind, motor-fuel-grade ethanol plant that produces a nominal 100 million US gallons per year, generates the dryhouse material balance illustrated in Table 3. The stillage volumes and solids concentrations shown are representative of a plant producing a nominal 14% v/v ethanol beer and recycling approximately 30% of the thin stillage as dilution water (backset) in grain mashing. The evaporator condensate and dryer vent flow rates indicate the energy demand of the corresponding unit operations. Solids concentrations in the evaporator condensate and dryer vent are due to the presence of volatile organic compounds in the stillage. operates with low maintenance and is robust while delivering high suspended-solids capture rates in addition to a high total-solids wet cake.

Centrifugation
Continuous decanter or solid-bowl centrifuges are generally used for whole-stillage solids separation, clarifying the centrate as well as dewatering the fibrous wet cake. The eight major components of the decanter centrifuge are shown in Figure 3:

SOLIDS SEPARATION
Dryhouse operations begin with partitioning the whole stillage suspension into fibre- and suspended-solidsrich wet cake and dissolved-solids-rich, thin-stillage fractions. A number of suspended-solids separation technologies have been used with varying degrees of success. These technologies include filtration systems that use inclined wedge wire or vibrating screens, or various types of belt or leaf filter presses. With wholestillage substrates, filtration produces a low-solids cake as well as a filtrate with high suspended solids. Postfiltration presses improve the concentration of cake solids, but high sheer increases the concentration of suspended solids in the pressate. Applying organic polymers improves solids separation, but this process is generally not cost effective. In the separation of liquid solids, the centrifugal force generated by the rotating assembly replaces the weaker force of gravity. High gravitational forces generated at very high rotational speeds efficiently separate suspended solids across a range of particle sizes in addition to dewatering the wet cake. The optimum technology is one that is capable of processing high hydraulic flow rates, requires moderate capital costs,
Whole stillage Solids, tph1 Water, tph1
1

Figure 3. Decanter centrifuge components

1. Solid bowl, the primary rotating assembly that contains the fluid and applies gravitational force for separating the suspended solids. 2. Main drive, which provides the rotational energy to the solid bowl, producing gravitational force. 3. Scroll conveyor, the internal rotating assembly that conveys the suspended-solids-rich cake from the rotating solid bowl. 4. Differential back drive, which provides breaking energy that reduces the speed of the scroll conveyor, producing a conveying effect and scraping the suspended solids along the internal wall, up the beach and out of the solids discharge port of the solid bowl. 5. Feed zone, which introduces the feed into the rotating solid bowl with minimal turbulence. 6. Beach, the inclined section of the solid bowl that extends above the pool of liquid and provides an area where gravitational force compresses and partially dewaters the suspended-solids-rich cake. 7. Solids discharge ports, the discharge points for suspended-solids-rich cake near the top of the beach.
Evaporator feed 17.14 178.56 Evaporator condensate 0.07 145.76 Evaporator syrup 17.07 32.80 DDGS 34.52 4.25 Dryer vent 0.02 62.14

Table 3. Dryhouse mass balance (for a dry-grind facility producing 100 million US gallons per year). Thin stillage 23.64 246.30 Wet cake 17.48 33.59 Backset 6.50 67.73

41.12 279.88

Tons per hour.

Dryhouse technologies and DDGS production 307 8. Filtrate ports with adjustable weirs, which are openings in the end of the solid bowl opposite the beach that provide control of fluid pool depth, inventory, residence time and a discharge point for clarified liquid. The design of the decanter centrifuge, in conjunction with centrifuge settings, determines the settling velocity of particles as well as the particle critical diameter, which together affect the capture efficiency of the suspended solids. For whole-stillage solutions that contain suspended solids with a wide range of particle sizes, the settling velocity is defined by Stokess law (Figure 4), where ut = settling velocity 2 = rate of rotation (radians per second) r = radial distance p = solid density = liquid density Dp = particle diameter = fluid viscosity Q = flow rate = liquid viscosity s = liquid depth p = solid density = liquid density V = liquid volume = rate of rotation (radians per second) r = radial distance

Dpc =

9Qs (p)V2r

Figure 5. Cut point equation for determining particle critical diameter.

Decanter centrifuges provide three basic controls to the plant operator: Feed rate, which adjusts the effective residence time Weir height, which controls the liquid volume and pool depth Back-drive torque, which adjusts the differential speed between the solid bowl and the scroll conveyor The effects of back-drive torque adjustments on cake solids and centrate clarity are illustrated in the following figures. In Figure 6, increasing torque by reducing the solid-bowl and scroll-conveyor differential speed results in an increased suspendedsolids residence time in the dewatering beach as well as in higher wet-cake total solids. Figure 7 illustrates the capture efficiency of suspended solids as a function of wet-cake total solids. Together, Figures 6 and 7 show

ut =

2r(p)Dp2 18

Figure 4. Stokess law equation for determining settling velocity.

With a given decanter centrifuge and whole-stillage flow rate, particle critical diameter, or cut point, is defined by Figure 5, where Dpc = particle critical diameter, or the smallest particle diameter capable of being separated,
50

Wet cake total solids (% wt)

45

40

35

30

25 0.0

1.0

2.0

3.0

4.0

5.0

Back-drive torque (kilonewtons per metre)

Figure 6. Effects of back-drive torque adjustment of a decanter centrifuge on suspended solids.

308 D.A. Monceaux and D. Kuehner


45 43 41 Wet cake solids (% wt) 39 37 35 33 31 29 27 25 85 86 87 88 89 90 91 92 93 94

Suspended solids capture efficiency (%)

Figure 7. Cake dryness versus the capture efficiency (recovery) of suspended solids.

that centrifugation is a compromise between cake solids and centrate clarity.

Radiation, which is heat transfer by electromagnetic radiation Dryhouse thermal processes remove water from various stillage streams, increasing whole-stillage solids from approximately 15% w/v total solids and producing a 90% total solids DDGS coproduct. The dryhouse of the nominal 100-million-US gallon-per-year facility represented in Table 3 requires the removal of 218 tons of water per hour. Figure 8 presents the enthalpy (the quantity of heat contained in one kilogram of water at the selected temperature), or energy available in water as a saturated liquid and vapour, and the associated temperature at various pressures. The ability to use water as an energy-transport mechanism is critical to efficient dryhouse operations.
500 400 350 300 250 200 150 100 50 132.26 191.05 264.53 338.01 411.49 0 F Enthalpy liquid (BTU/lb) 450

Thermal processes
High-flow aqueous process solutions are found throughout the dry-grind fuel ethanol plant. Extensive movement of energy from streams needing to be heated or cooled in the various unit operations is needed occurring by one of the following mechanisms: Conduction, which is heat transfer by means of molecular agitation within a material without any motion of the material as a whole Convection, which is heat transfer by mass motion of a fluid such as air or water when the heated fluid is caused to move away from the source of heat, carrying energy with it
1220 1200 Enthalpy vapour (BTU/lb) 1180 1160 1140 1120 1100 1080 1060 1040 11.76 19.10 24.98 35.27 58.78 1020 Enthalpy vapour deg F Enthalpy liquid

Steam pressure (psia)

Figure 8. Water enthalpy chart.

95.52

0.29

1.03

4.41

Dryhouse technologies and DDGS production 309

Evaporation
Evaporation technology uses convective heat transfer to concentrate nonvolatile substances in solution or suspension, producing higher-solids-concentration products. The energy-driving evaporation process uses steam or other process streams. In evaporation, energy is applied to a liquid at constant pressure, raising the temperature to saturation the point where it holds as much energy as possible without boiling. As additional energy is applied, the vapour pressure of the liquid reaches the vapour pressure of the surrounding environment, and the liquid begins to vaporise. The heat of vaporisation is the amount of energy required for the liquid to change state to a vapour without a change in temperature. The resulting vapour separates from the residual liquid, increasing the concentration of the nonvolatile fraction. The heat-transfer process is defined by Fouriers law (Figure 9), where Q = the rate of heat to be transferred UO = the overall coefficient of heat transfer AO = the area of heat-transfer surface Tlm = the logarithmic mean temperature difference

and nondistillable fermentation end products. During stillage evaporation, thin stillage, which contains from 5 to 10% total solids, is concentrated to produce a nominal 30 to 50% total solids. Figure 10 illustrates the temperature, enthalpy and state relationship of water, showing the quantity of energy required and available as water changes from a liquid to a vapour during the evaporation process.

Evaporator systems
A simple industrial evaporator system (Figure 11) will contain the following: Calandria, or heat exchanger, which transfers energy from the source stream to the solidscontaining fluid, raising the fluids temperature to the boiling point Circulation, or feed, pump, which supplies feed to the evaporators heat exchanger Distributor, which distributes feed or circulating fluid evenly across the faces of the tube sheets of the tubular evaporator calandrias, ensuring that the surfaces of the gravity-fed tubes are thoroughly wetted Transfer pump, which moves enriched-solidscontaining fluid from the evaporator calandria Vapour separator, which separates the water vapour from the enriched-solids-containing fluid Condenser, which removes energy from the evaporator via heat transfer with another fluid Vacuum source, which removes noncondensable components in the vapour
Heating system

Q = U AOTlm O

Figure 9. Fouriers law equation defining the heat-transfer process.

Thin stillage is predominantly an aqueous suspension of soluble and insoluble grain solids

212

32

Temperature (F)
144 BTU/lb to melt ice 180 BTU/lb to at 32F heat water from heat of 32F to 212F fusion

972 BTU/lb to vaporise water at 212F heat of vaporisation

Heating ice

Energy addition

Figure 10. The relationship of water temperature, enthalpy and the state of water.

310 D.A. Monceaux and D. Kuehner improvement in system efficiency. In these designs, the vapour directed to the condenser (shown in the simple evaporator illustrated in Figure 11) is routed to succeeding evaporation stages, or effects. So, the first evaporator effect condenses the incoming steam, producing a near-equal amount of vapour that condenses in the second effect.

Figure 11. Components of a simple evaporator.

The design of a simple evaporator is shown in Figure 11, where approximately one unit of steam is condensed on the shell of the calandria, transferring the heat of condensation to feed located in the tubes, evaporating one unit of water. When the feed temperature is below saturation, additional energy will be required to raise the liquid to the boiling point at the system pressure. The vapour produced in the evaporator flows through the separator, removing entrained liquid before condensing and transferring the energy to cooling water, with the concentrated product pumped to storage. The uncontaminated steam condensate is returned to the boiler for reuse. Together, the components are referred to as an evaporator effect. For energy transfer to occur, a temperature differential must exist across the heat-transfer surface area. Because steam and stillage are both predominately water, a temperature differential must be accompanied by a corresponding pressure differential, as illustrated in Figure 8. The efficiency of the simple evaporator system in Figure 12 results in about one unit of steam removing one unit of water with a near-equal quantity of energy transferred to the cooling water. Figures 13 through 15 illustrate options for further

Figure 12. Single-effect evaporator.

As seen in these figures, additional evaporator effects remove more water per unit of steam supplied, and system efficiency improves. From this, it appears that evaporation systems could be infinitely efficient through the addition of effects, but design and operating parameters dictate otherwise. Critical issues include the minimal practical condensing temperature of the final vapour, which is a function of the cooling water temperature, the maximum practical product side temperature in the first effect, which is a function of the thermal stability and fouling potential of the feed, the practical differential temperature across the individual effects, considering operating parameters such as fouling of heat-transfer surface area and boiling-point elevation of the product and cleaning frequency.

Dryhouse technologies and DDGS production 311

Figure 13. Two-effect evaporator.

Figure 14. Three-effect evaporator.

312 D.A. Monceaux and D. Kuehner

Figure 15. Four-effect evaporator.

A multi-effect stillage evaporation system, with a firsteffect product side temperature of 210F and a final condensate temperature of 130F, that operates with a 15F differential temperature across each effect could be designed with approximately five effects.

Thermocompression evaporators
Alternatives exist to improve evaporator efficiency without the continued, capital-intensive addition of effects. Such evaporator systems have been used in dryhouse designs for over 25 years. These systems increase efficiency by recycling vapour from later effects to preceding effects in the evaporator. To accomplish this, the pressure of the vapour must be increased by thermocompression (Figure 16) to offset the design-basis pressure drop of the system. Evaporator vapour is generally boosted by taking a portion of the vapour from one of the effects and directing it to a steam ejector (Figure 17). This device produces a Venturi effect, where ejector fluid under high pressure is converted into a high-velocity jet at the throat of the nozzle, which creates a low pressure at that point. The low pressure draws the suction fluid

into the nozzle, where it mixes with the motive fluid, resulting in an intermediate pressurevapour mixture. The quantity of vapour recycled is a function of the design of the ejector, the motive steam pressure and the pressure of the evaporator vapour. A disadvantage of steam ejector systems is that the motive steam is often contaminated with impurities present in the evaporator vapour. In stillage evaporators, the condensate contains measurable concentrations of ethanol and organic acids and cannot be reused as boiler feed makeup water.

Mechanical compression evaporators


Another variant of vapour-compression technology uses electrical or steam-turbine-driven devices such as fans, blowers or compressors to boost the pressure and recycle the evaporator vapour. In mechanical vapour recompression (MVR) evaporation (Figure 18), the vapour from the separator, free of entrained liquid, is compressed, elevating the condensing temperature. The vapour directed to the shell of the evaporator body condenses, transferring energy back to the circulating fluid. A small amount of

Dryhouse technologies and DDGS production 313

Figure 16. Thermocompression evaporator.

additional energy is required to balance the systems enthalpy, replacing energy that is required to raise the temperature of the incoming feed to the operating conditions of the evaporator.

exchanger surface area (higher boost pressures increase vapour temperatures, providing greater differential temperatures across evaporator bodies, reducing heat-transfer surface areas). In stillage evaporation, compressing large vapour flows translates to high power demand. MVR systems are best suited for applications where high-pressure steam is available for an exhaust, an extractive turbine-drive is applied and low-cost power allows for an electrical drive.

Figure 17. Steam ejector (Croll Reynolds Company, Inc.).

Fluid properties and evaporator system design


Fluid physical properties must be taken into consideration during the design of evaporator systems. Stillage is a complex mixture of inorganic salts, organic acids, soluble and insoluble proteins, peptides and amino acids, carbohydrates, sugar alcohols such as glycerol, lipids and fibre fines.

Design-basis considerations for MVR systems are a compromise between compressor power (increasing the pressure differential increases power demand, resulting in lower system reliability) and

314 D.A. Monceaux and D. Kuehner effective differential temperature and increases the required heat-transfer surface area. Another physical property having a significant impact on evaporator design and performance is fluid viscosity. As dilute feed streams concentrate, fluid viscosities increase, exhibiting both Newtonian and non-Newtonian properties: A Newtonian fluid is a fluid whose viscosity does not change with the rate of flow or shear stress. A non-Newtonian fluid is a fluid whose viscosity changes with the rate of flow or shear stress. Figure 19 illustrates the sheartemperature relationship in thick-stillage streams that contain high concentrations of suspended solids. Elevated fluid viscosities interfere with film formation in falling-film evaporation systems, resulting in uneven wetting of heat-transfer surfaces. In high-velocity forced circulation evaporation systems such as finishing evaporators, increasing viscosities generate laminar flow, reduced Reynolds numbers and corresponding lower heat-transfer coefficients. Increasing fluid velocity in an effort to improve heat transfer has the adverse effect of increasing system pressure drop and power demand. In all instances, increased fluid viscosities accelerate fouling of heat-transfer surfaces, with fouling being most severe in high solids effects.

Figure 18. Mechanical vapour recompression (MVR) evaporator.

A key property is fluid boiling-point elevation. When a solute is added to a solvent, the vapour pressure of the solvent (above the resulting solution) is less than the vapour pressure above the pure solvent. The resulting boiling point of the solution will be greater than the boiling point of the pure solvent. This is because the solution (which has a lower vapour pressure) must be heated to a higher temperature in order for the vapour pressure to become equal to the external pressure. During evaporation, as the solids concentration of the stillage increases, boiling-point elevation reduces the
200 180 160 Centipoise x 1,000 140 120 100 80 60 40 20 0 0.3 0.6 1.5

Evaporator clean-in-place (CIP) systems


The design-basis criteria of a stillage evaporation system must consider fouling rate and the associated cleaning process. Most evaporation systems are equipped with CIP systems that chemically
43 C 54 C 65 C

RPM

12

Figure 19. Concentrated stillage viscosity profile over a range of temperature and revolutions per minute during testing.

Dryhouse technologies and DDGS production 315 clean heat exchangers, separators and associated product-side piping without opening the equipment. Heat-transfer-surface fouling rate and the associated CIP frequency and CIP cycle time detract from the onstream time of the evaporator system. In heavily fouled stillage evaporators, one finds complex, mixed organic and inorganic deposits on the heat-transfer surface. Cleaning stillage evaporators requires the following steps, which can require up to twenty hours to complete: Initial rinse. Hot water flushes stillage from the system and rinses away easily removable deposits, reducing chemical consumption during subsequent CIP steps. Caustic wash. Dilute caustic solution attacks and partially solubilises the deposits organic matrix. Intermediate rinse. Hot water rinses dilute caustic from the system in advance of the ensuing acid wash. Acid wash. A dilution acid solution, such as sulfamic acid, attacks and partially solubilises the calcium oxalate and calcium-sulphate-rich inorganic matrix of deposits. Final rinse. Hot water flushes residual acid and dislodged deposits from the system. and drying systems have many functional similarities, including an energy source, mechanisms for introducing feed into the drying system, a conditioning system to ensure that feed and product flow freely in the dryer, heat-transfer mechanisms and vapourproduct separation equipment. In addition to the thermodynamic principles of Fouriers law such as heat duty, heat-transfer rate and temperature differentials, dryer design and operations must also consider three interrelated factors that impact dryer selection and operations: particle residence time, temperature sensitivity of the product and bound moisture. The presence of bound, or encapsulated, moisture (Figure 20) the water that is chemically bound to cellulose, hemicellulose, lignin or similar compounds and is difficult to remove increases the residence time in the dryer. In many cases temperature must also be increased, adversely affecting the quality of temperature-sensitive products.

Solids drying
Like evaporation, drying is a mass-transfer process resulting in the removal of water or moisture from a process stream. While evaporation increases the concentration of nonvolatile components in solution, in drying processes the final product is a solid. Drying processes reduce the solute or moisture level to improve the storage and handling characteristics of the product, maintain product quality during storage and transportation and reduce freight cost (less water to ship). Industrial drying applications use conductive and/ or convective heat-transfer processes to reduce the concentration of residual volatile components in process streams that are rich in nonvolatile compounds. The principles of solids drying are similar to those of other thermal processes such as evaporation. Consequently, industrial evaporators
Figure 20. Encapsulated, or bound, moisture.

Dryer categories and selection


Table 4 categorises continuous industrial dryer technologies by their methods of heat transfer and product conveyance. Several of these technologies have been used in the dry-grind ethanol industry, with varying degrees of success. Dryer technology selection criteria includes a combination of external factors and drying systems issues (Table 5), several of which are common to both categories.

316 D.A. Monceaux and D. Kuehner


Table 4. Continuous dryer technologies. Heat-transfer mechanism Direct dryers Direct contact between wet solids and hot gases Vaporised liquid conveyance By the drying media Material conveyance Mechanical and/or pneumatic Dryer types Tray, sheet, rotary*, tunnel, through circulation, SSD** rotary Pneumatic conveyor*, ring*, fluid-bed, spray, SSD** ring Cylinder, drum, steambelt, screw conveyor, steam tube*

Pneumatic by drying media Indirect dryers To wet solids through a retaining wall Independently from heating media Mechanical

* Application in dry-grind ethanol production. ** SSD: Super heat steam dryer. Table 5. Comparison of dryer issues. External factors Plant location Local rules and regulations Energy source Emission requirements DDGS market Upstream processes Energy recovery Drying system issues Reliability and operability Energy efficiency Energy source Emission requirements Product quality Capital investment Energy recovery

Dryer exhaust
Increasing fuel prices have led the ethanol industry to pursue strategies that reduce the net energy cost of producing DDGS. This has resulted in improvements in dryer efficiency as well as in combustion systems capable of using lower-priced fuels such as coal, biomass and forest products. In addition, process technology providers continue to pursue opportunities to utilise the dryer exhaust as a source of energy in other ethanol plant operations, reducing overall plant energy consumption. The properties of the dryer exhaust are a function of the dryer technology used and affect the ability to recover the energy value of the dryer. The ability to effectively recover energy contained in dryer exhaust is primarily a function of the dew point of the vapour. Dew point is defined as the temperature at which water will begin to condense from the exhaust under constant pressure. In dryer systems, the dew point of the exhaust stream is influenced by its composition. Aside from water vapour, the dryer exhaust is composed of air, products of fuel combustion, particulate matter and volatile compounds present in or resulting from the thermal decomposition of the dryer feed. The concentrations of the various nonwater components are a function

of dryer design, maintenance and operations. Air is generally introduced into dryers during the fuel combustion process, as a sweep gas to assist in conveying moisture and solids or as a result of leakage. Dryer design selection is impacted by air moisture concentration and associated dew point, as illustrated in Figure 21. As the fuel ethanol industry has grown, it has attracted the attention of environmental regulatory agencies. During the 1990s it was determined that DDGS dryer exhaust was a major source of priority pollutants. This included volatile organic compounds such as acetic acid, ethanol and furfural, as well as particulate matter and products of the drying process. In addition, as the ethanol production capacity of plants increased, products of combustion such as nitrogen oxides, sulphur oxides and carbon monoxide became a factor in permitting new plants. The industry and technology suppliers responded with the addition of emissions control devices like thermal oxidisers to conventional dryer technologies as well as the development of new dryer technologies. Recovering dryer energy in other processes results in the condensation of substantial amounts of water from the exhaust stream. This has the effect of reducing the organic load to and thermal duty on the thermal oxidisers at the expense of generating a high biochemical oxygen demand (BOD) in the wastewater stream.

DDGS drying systems


Various dryer technologies have been used during the course of the development of the dry-grind ethanol industry. Early in the industrys history, most dryhouses were more or less based upon beverage

Dryhouse technologies and DDGS production 317


Partial gas recycle and steam tube dryer

220 210 200 190 180 170 160 150 140 130 120 110 100 90 80 70 60 50 40 30

Dew point (temperature F)

Superheated steam rotary drum dryer

Superheated steam ring dryer

Open circuit dryer

6 8 Water (lbs)/dry air (lbs)

10

12

Figure 21. Impact of exhaust dew point and moisture content in relation to dryer design.

distillery industry standards, with DDGS drying operations that predominantly used steam-tube dryers. As the industry matured, average plant capacity increased and DDGS transformed from a by-product to a value-added source for animal nutrition. Rotary and ring-dryer technologies began to displace steamdriven systems. Of the numerous options that have been implemented during the past three decades, three basic configurations dominate the industry today steam-tube dryers, rotary dryers and ring dryers. The fundamental design for a DDGS dryer incorporates the following basic process steps: Furnaces combust fuels, generating hot gases that are used directly or indirectly (SSD) as a source of heat for drying. In the case of steam-tube dryers, the energy source lies in the boiler, generally independent of the dryer. Solids handling system and pumps continuously feed, convey and discharge wet cake, solubles and DDGS. Feed conditioning and mixers blend a portion of the dry DDGS product and wet incoming feed, changing the physical properties and handling characteristics of the feed streams and reducing the agglomeration of solids and plugging of dryer internals. Dryer body moves solids in rotating flighted drums or vertical ducts to contact the solids and hot gas streams or surfaces. Product recovery separators and cyclones remove DDGS solids and fine particulate matter from the gas/vapour streams.

Product coolers reduce the temperature of the dry DDGS product to near-ambient temperature, improving product handling and reducing the opportunity of spontaneous combustion during storage and transit. Emissions-control cyclones, scrubbers and thermal oxidisers reduce the emission of particulate matter, carbon monoxide and volatile organic compounds (VOCs). Air handling system moves vapour and hot gases during the drying process. Various approaches that use these systems are compared in Table 6 and illustrated in the figures that follow.

Rotary direct-fired PGR dryers


To date, rotary direct-fired PGR dryers (RDFDs), shown in Figure 22, have the greatest market penetration in North America. This is primarily due to a significant base that was installed before increased attention from regulatory agencies and the mandated application of thermal oxidation systems. Of the four dryer technologies illustrated, a lower construction cost has kept the RDFD in high demand. The lower cost is a result of the dryers ability to operate at higher inlet gas temperatures. Since higher inlet gas temperatures reduce the time that solids must reside in the dryer body, the size of the dryers rotating drum can be reduced, with a comparable reduction in equipment cost. However, higher temperatures result in reduced product quality, increased VOC emissions and higher equipment maintenance. Modern RDFDs are equipped

318 D.A. Monceaux and D. Kuehner


Table 6. Characteristics of DDGS dryer technologies. Rotary direct-fired PGR3 (RDFD) Figure Type Heat source Heat exchange Solids transport 22 Direct-fired Combustion gas Direct hot gas solids contact Mechanical/pneumatic in rotating drum Ring PGR* 23 Direct-fired Combustion gas Direct hot gas solids contact Pneumatic in highvelocity gas recycle Rotary indirect-fired SSD (RIFD)/Ring SSD 24 Indirect-fired Combustion gas Direct superheated steam solids contact Mechanical/pneumatic in rotating drum or pneumatic in highvelocity gas recycle Internal purge gas directed to furnace Heat exchanger between combustion gas and drying media Rotary steam-tube dryer (TSTD) 25 Indirect Steam Hot tube wall solids contact Mechanical in in rotating drum

Emission control Miscellaneous

EOS TO***

EOS TO*** Disintegrator required to control particulate size of solids

EOS TO*** Heat-transfer surface required in dryer and boiler

* PGR: Partial gas recycle. ** SSD: Superheated steam dryer. *** EOS TO: End-of-stack thermal oxidiser.

Figure 22. Rotary direct-fired PGR dryer.

with partial recycle of the exhaust (partial gas recycle), resulting in reduced fire risk (lower oxygen content) as well as improved energy efficiency and reduced dryer emissions. End-of-pipe thermal oxidation systems are generally used for emissions control. Alternatively,

some RDFD installations have integrated the dryer exhaust thermal oxidisation process with waste-heat steam generation to improve upon overall plant energy balance. Unfortunately, this has linked dryer emissions treatment with steam generation for the

Dryhouse technologies and DDGS production 319 ethanol plant, but these two processes do not always operate synchronously. while retaining heavier and larger particles. End-ofpipe thermal oxidation systems are generally used for emissions control. Due to low air infiltration, the size and operating cost of the TO is reduced.

Ring dryers
Ring dryers (RDs) follow rotary direct-fired dryers in market penetration (Figure 23). Compared to rotary direct-fired dryers, these systems show similar installed capital investment but improved primary energy efficiency and product quality. When compared to a rotary dryer, pneumatic transport of the product in the ring dryer body increases electrical energy consumption. The primary energy efficiency of an RD is due, to a large extent, to the high hot-gasrecycle rate and well-sealed design. These features result in low air entrainment, producing a high dew point exhaust gas and offering greater opportunity for waste-heat recovery applications. The design reduces the time that the DDGS solids are subjected to heat, improving product quality. This is especially important when processing high-protein feeds. The short residence time is possible because of the application of separation, classification and particle-size reduction technologies. Combined, these serve to control particle size, selectively removing dry product from the system

Indirect-fired SSD dryers


Indirect-fired SSD dryers (Figure 24a) follow RDFDs and RDs in market share. The system employs full gas recycle, but instead of introducing hot combustion gases directly into exhaust recycle, energy is applied indirectly via a heat exchanger. The exchanger transfers heat from the furnace combustion gases to the recirculating exhaust, superheating the stream. As the superheated dryer exhaust is reintroduced into the dryer, the energy is transferred to the product, vaporising water without condensing. For this reason, the technology is often referred to as superheated steam drying, or SSD. One major feature of the indirect-fired SSD dryer is the integration of emissions control technology provided by operating the furnace under conditions suitable for thermal oxidation of organic compounds in the dryer purge. In addition, the design of the closed SSD loop increases the purge gas energy recovery potential, providing even greater

Figure 23. Ring PGR dryer.

320 D.A. Monceaux and D. Kuehner

Figure 24a. Indirect-fired SSD dryer rotary.

Figure 24b. Indirect-fired SSD dryer ring.

Dryhouse technologies and DDGS production 321 opportunity for waste-heat recovery applications. Due to reduced air entrainment, the ring SSD (Figure 24b) produces the highest dew point, and, therefore, the highest energy recovery potential. Furthermore, ring SSD systems can be pressurised, which increases energy recovery potential. Because rotating drums are difficult to completely seal, air infiltration is increased, and the exhaust dew point is reduced. The installed cost of indirect-fired SSDs is comparable with that of similar-sized RDFDs or RDs complete with thermal oxidiser systems. Reduced operating cost when running the indirect-fired SSD with energy recovery often results in better economics. demand per unit of water evaporated in RSTDs is generally higher than that of other dryer technologies. Likewise, capital investment increases due to the considerable heat-transfer surface area in the dryer and the associated steam generation system. Extended contact of the product with the surface of the steam-containing tubes results in protein denaturation and reduced product quality. RSTDs can be designed with relatively low air infiltration rates, but because the rotating drum is not as well-sealed as a ring dryers duct, air infiltration is increased and the exhaust dew point is reduced.

Rotary steam-tube dryers


Rotary steam-tube dryers (RSTD) continue to be used where applications warrant (Figure 25). These include circumstances where appropriately priced steam is available, where fuel selection does not allow for direct-fired applications or where the fuel is incompatible with the use of indirect-fired SSD dryer heat exchangers. Major impediments to greater market penetration include low energy efficiency, high capital cost and reduced product quality. In the case of RSTDs, energy consumption is a function of both dryer and boiler efficiency. Consequently, energy

DRYHOUSE OPERATIONAL ISSUES


Major unit operations comprising the dryhouse incorporate numerous control points to enable plant personnel to optimise operations. Aside from the designated control mechanisms, numerous external parameters can have a major impact on dryhouse operations. These parameters primarily relate to two general categories of process fluid physical properties: particle size distribution and fluid viscosity. These properties are, to a degree, controllable by plant design and/or operator decisions.

Figure 25. Rotary steam-tube dryer.

322 D.A. Monceaux and D. Kuehner Solids distribution studies reveal that fine-grind milling results in high concentrations of fine particulate matter in process streams. This serves to reduce decanter centrifuge separation efficiency, increases non-Newtonian viscosity of thin stillage and stillage concentrates and adversely impacts DDGS dryer operations by disrupting wet cake, syrup and DDGS mixing. Dryhouse operations are further compromised by high fermentation and stillage acidity concentrations. The presence of elevated populations of acid-producing microbes or the use of excessive quantities of mineral acids increases process-fluid acidity and degrades fibre structure, producing fibre fines that reduce the separation efficiency of decanter centrifuge solids. Because fibre destruction is a function of acid concentration, temperature and time, technology providers must be concerned with beer-still operation temperatures as well as with whole-stillage storage conditions. In addition to non-Newtonian viscosity generated by the presence of suspended solids, numerous front-end processes can serve as sources of Newtonian viscosity. Sources of non-Newtonian viscosity include unconverted and retrograded starch resulting from inefficient cooking and liquefaction operations, unfermented dextrin and glucose concentrations resulting from poor temperature control and microbial activity, elevated lactic acid concentrations resulting from microbial contaminant activity and elevated glycerol concentrations resulting from environmental stress on yeast. As the concentration of undesirable solids in the stillage increases during dryhouse operations, so does fluid viscosity. These additional solids not only detract from ethanol yield but also increase dryhouse duty and associated energy consumption. Therefore, optimisation of dryhouse operations is greatly dependent on nondryhouse processes. ethanol production has focused more attention on the industry. Opportunities are being explored to improve production economics, focus research activities and develop traditionally undervalued stillage streams. These efforts include the application of dry-grind corn fractionation technologies that recover higher value protein and lipid-rich concentrates in advance of the ethanol process. This reduces mass flow through the ethanol process as well as solids loading on dryhouse operations. In addition, groups continue to research stillage-refining technologies that promise to recover and concentrate higher-value constituents of whole stillage such as glycerol, organic acids, amino acids, peptides, proteins and lipids. The application of these new technologies, individually or in various combinations, will change the technical definition of DDGS as well as the unit operations and processes used in their production. As the industry continues to grow, the dryhouses of the past will not be easy to find in the ethanol production facilities of the future.

REFERENCES
Anonymous (2008). College of Food, Agricultural and Natural Resource Sciences, University of Minnesota, St. Paul, MN. Belyea RL, Rausch KD and Tumbleson ME (2004) Composition of corn and distillers dried grains with solubles from dry grind ethanol processing. Bioresource Technology 94 293-298. Croll Reynolds (2008) Process Vacuum and Power Systems Thermocompressor Theory. http://www. croll.com/_website/pr/thermhome.asp (accessed April 2008). Green DW and Perry RH (2007) Perrys Engineering Handbook 8th ed. McGraw-Hill, New York, NY. Minton PE (1986) Handbook of Evaporation Technology. Noyes Publications, USA. Rausch KD and Belyea RL (2005) Coproducts from bioprocessing of corn. ASAE Annual International Mtg., Tampa, Fl. July 17-20. Paper No. 057041. Shurman J (2008) Distillers Grains By-products in Livestock and Poultry Feeds. University of Minnesota, Department of Animal Science. http:// www.ddgs.umn.edu/profiles.htm (accessed 10 April 2008). Woon-Feung Leung Wallace (1998) Industrial Centrifugation Technology 1st ed. McGraw-Hill, New York, NY.

CONCLUSION
The dryhouses found in todays dry-grind fuel ethanol facilities are not significantly different from those found in early whisky distilleries. Decanter centrifuges, thin-stillage evaporators and thermal dryers continue to be widely used to convert whole stillage to DDGS. Recent growth in grain-based fuel

Você também pode gostar