Sei sulla pagina 1di 25

The heating of liquids in tanks is an important requirement in process industries such as the dairy, metal treatment and textile

industries. Water may need to be heated to provide a hot water utility; alternatively, a liquid may need to be heated as part of the production process itself, whether or not a chemical reaction is involved. Such processes may include boiler feedtanks, wash tanks, evaporators, boiling pans, coppers, calandrias and reboilers. Tanks are often used for heating processes, of which there are two major categories:

Totally enclosed tanks, such as those used for storing fuel oil, and where heat load calculations are generally straightforward. Open topped tanks, where heat load calculations may be complicated by the introduction of articles and materials, or by evaporative losses.

Open and closed tanks are used for a large number of process applications:

Boiler feedtanks - The boiler feedtank is at the heart of any steam generation system. It provides a reservoir of returned condensate and treated make-up water, for feeding the boiler. One reason for heating the water is to reduce oxygen entering the boiler, with (theoretically) 0 ppm oxygen at 100C. Boiler feedtanks are normally operated at between 80C and 90C. Hot water tanks - Hot water is required for a number of processes in industry. It is often heated in simple, open or closed tanks which use steam as the heating medium. The operating temperature can be anywhere between 40C and 85C depending on the application.

Degreasing tanks - Degreasing is the process where deposits of grease and cooling oil are removed from metal surfaces, after machining and prior to the final assembly of the product. In a degreasing tank, the material is dipped into a solution, which is heated by coils to a temperature of between 90C and 95C.

Metal treatment tanks - Metal treatment tanks, which are sometimes called vats, are used in a number of different processes:

To remove scale or rust.

To apply a metallic coating to surfaces.

The treatment temperatures typically range from 70C to 85C.

Oil storage tanks - Storage tanks are required to hold oils which cannot be pumped at ambient temperatures, such as heavy fuel oil for boilers. At ambient temperatures, heavy oil is very thick and must be heated to 30C - 40C in order to reduce its viscosity and allow it to be pumped. This means that all heavy oil storage tanks need to be provided with heating to facilitate pumping. Heating tanks used in process industries - Heating tanks are used by a number of process industries, see Table 2.9.1.

Tabl e 2.9.1 Process industries which use heating tanks In some applications the process fluid may have achieved its working temperature, and the only heat requirement may be due to losses from the solid surface of the walls and/or the losses from the liquid surface. This Tutorial will deal with the calculations which determine the energy requirements of tanks: the following two Tutorials (2.10 and 2.11) will deal with how this energy may be provided. When determining the heat requirement of a tank or vat of process fluid, the total heat requirement may consist of some or all of a number of key components:

The heat required to raise the process fluid temperature from cold to its operating temperature. The heat required to raise the vessel material from cold to its operating temperature. The heat lost from the solid surface of the vessel to the atmosphere. The heat lost from the liquid surface exposed to the atmosphere. The heat absorbed by any cold articles dipped into the process fluid.

However, in many applications only some of the above components will be significant. For example, in the case of a totally enclosed well-insulated bulk oil storage tank, the total heat requirement may be made up almost entirely of the heat required to raise the temperature of the fluid.

Items 1 and 2, the energy required to raise the temperature of the liquid and the vessel material, and item 5, the heat absorbed by any cold articles dipped into the process fluid, can be found by using the Equation 2.6.1. Generally, data can be accurately defined, and hence the calculation of the heat requirement is straightforward and precise.

Equation 2.6.1 Items 3 and 4, the heat losses from the vessel and liquid surfaces can be determined by using Equation 2.5.3. However, heat loss calculations are much more complex, and usually empirical data, or tables based on several assumptions have to be relied upon. It follows that heat loss calculations are less accurate. Equation 2.5.3

Heat loss from the solid surface of the vessel to the atmosphere
Heat will only be transferred provided there is a difference in temperature between the surface and the ambient air. Figure 2.9.1 provides some typical overall heat transfer coefficients for heat transfer from bare steel flat surfaces to ambient air. If the bottom of the tank is not exposed to ambient air, but is positioned flat on the ground, it is usual to consider this component of the heat loss to be negligible, and it may safely be ignored.

For 25 mm of insulation, the U value should be multiplied by a factor of 0.2. For 50 mm of insulation, the U value should be multiplied by a factor of 0.1.

The overall heat transfer coefficients provided in Figure 2.9.1 are for 'still air' conditions only.

Fig. 2.9.1 Typical overall heat transfer coefficients from flat steel surfaces Table 2.9.2 shows multiplication factors which need to be applied to these values if an air velocity is being taken into account. However, if the surface is well insulated, the air velocity is not likely to increase the heat loss by more than 10% even in exposed conditions.

Table 2.9.2 Effect on heat transfer with air movement Velocities of less than 1 m/s can be considered as sheltered conditions, whilst 5 m/s may be thought of as a gentle breeze (about 3 on the Beaufort scale), 10 m/s a fresh breeze (Beaufort 5), and 16 m/s a moderate gale (Beaufort 7). For bulk oil storage tanks, the overall heat transfer coefficients quoted in Table 2.9.3 may be used.

Table 2.9.3 Overall heat transfer coefficients for oil tanks

Water tanks: heat loss from the water surface to the atmosphere

Figure 2.9.2 relates heat loss from a water surface to air velocity and surface temperature. In this chart 'still' air is considered to have a velocity of 1 m/s, tanks in sheltered positions outdoors consider velocities at about 4 m/s, whilst tanks in exposed positions outdoors are considered with velocities at about 8 m/s. This chart provides the heat loss in W/m rather than the units of the overall heat transfer coefficient of W/mC. This means that this value must be multiplied by the surface area to provide a rate of heat transfer, as the water to air temperature difference has already been taken into account. Heat losses from the water surface, as shown in Figure 2.9.2 are not significantly affected by the humidity of the air. The full range of humidities likely to be encountered in practice is covered by the thickness of the curve. However, the graph considers heat losses with an air temperature of 15.6C and 55% air humidity. Different conditions to these can be calculated from the Calculators on this website. To determine the heat loss from the chart, the water surface temperature must be selected from the top scale. A line should then be projected vertically downwards to the (bold) heat loss curve. For indoor tanks a line should be projected horizontally from the intersection to the left-hand scale. For outdoor tanks a horizontal line should be projected either left or right until it intersects the required location, either sheltered or exposed. A projection vertically downwards will then reveal the heat loss on the bottom scale. In most cases, the heat loss from the liquid surface is likely to be the most significant heat loss element. Where practical, heat loss can be limited by covering the liquid surface with a layer of polystyrene spheres which provide an insulating 'blanket'. Any solution to reduce heat losses becomes even more important when tanks are located outside in exposed positions as portrayed by the graph in Figure 2.9.2.

Fig. 2.9.2 Heat loss from water surfaces Example 2.9.1 For the tank shown in Figure 2.9.3, determine: Part 1. The mean heat transfer rate required during start-up. Part 2. The maximum heat transfer rate required during operation.

Fig. 2.9.3 The tank is unlagged and open topped and is situated on a concrete floor inside a factory. It is 3 m long by 3 m wide by 2 m high. Tank total surface area = 24 m (excluding base). Heat transfer coefficient from tank/air, U1 = 11 W/mC. The tank is /3 full of a weak acid solution (cp = 3.9 kJ/kgC) which has the same density as water (1 000 kg/m) The tank is fabricated from 15 mm mild steel plate. (Density = 7 850 kg/m, cp = 0.5 kJ/kgC)

The tank is used on alternate days, when the solution needs to be raised from the lowest considered ambient temperature of 8C to 60C in 2 hours, and remain at that temperature during the day. When the tank is up to temperature, a 500 kg steel article is to be dipped every 20 minutes without the tank overflowing. (cp = 0.5 kJ/kgC)
M (start-up)

Part 1 Determine the mean heat transfer rate required during start-up This is the sum of: A1. Heating the liquid
M (liquid)

A2. Heating the tank material

M (tank)

A3. Heat losses from the sides of the tank A4. Heat losses from the liquid surface Part 1.1 Heating the liquid
M (liquid)

M (sides)

M (surface)

Equation 2.6.1

Equation 2.5.3 Where: T is the mean temperature difference T M

Part 2 Determine the running load, that is the maximum heat transfer rate required during operation (operation)

In operating conditions, the liquid and tank (A1 and A2, from Part 1) are already up to operating temperature, so the heating components = 0. In operating conditions, the heat losses from the liquid and tank (A3 and A4, from Part 1) will be greater. This is because of the greater difference between the liquid and tank temperatures and the surroundings. Immersing the article in the liquid is clearly the objective of the process, so this heat load must be calculated and added to the running load heat losses.

Part 2.1 Heat losses from tank sides Equation 2.5.3 Where:

Equation 2.6.1

Note that the operational energy requirement (59 kW) is significantly less than the start-up energy requirement (367 kW). This is typical, and, where possible, the start-up period may be extended. This will have the effect of reducing the maximum energy flowrate and has the benefits of levelling demand on the boiler, and making less demand on the temperature control system. For tanks that are to operate continuously, it is often only necessary to calculate the operating requirements i.e. the Part 2 calculations.

Vessels can be heated in a number of different ways. This tutorial will deal with indirect heating. In these systems, the heat is transferred across a heat transfer surface. Options include:

Submerged steam coils - A widely used form of heat transfer involves the installation inside a tank of a steam coil immersed in a process fluid. Steam jackets - Steam circulates in the annular space between a jacket and the vessel walls, and heat is transferred through the wall of the vessel.

Submerged steam coils


The use of tank coils is particularly common in marine applications where cargoes of crude oil, edible oils, tallow and molasses are heated in deep tanks. Many of these liquids are difficult to handle at ambient temperatures due to their viscosity. Steam heated coils are used to raise the temperature of these liquids, lowering their viscosity so that they become easier to pump. Tank coils are also extensively used in electroplating and metal treatment. Electroplating involves passing articles through several process tanks so that metallic coatings can be deposited on to their surfaces. One of the first stages in this process is known as pickling, where materials such as steel and copper are treated by dipping them in tanks of acid or caustic solution to remove any scale or oxide (e.g. rust) which may have formed. Steam coil sizing Having determined the energy required (in Tutorial 2.9), and with knowledge of the steam pressure/temperature in the coil, the heat transfer surface may be determined using Equation 2.5.3: Equation 2.5.3 The heat transfer area calculated is equivalent to the surface area of the coil, and will enable an appropriate size and layout to be specified. Determining the 'U' value To calculate the heat transfer area, a value for the overall heat transfer coefficient, U, must be chosen. This will vary considerably with the thermal and transport properties of both fluids and a range of other conditions. On the product side of the coil a thermal boundary layer will exist in which there is a temperature gradient between the surface and the bulk fluid. If this temperature difference is relatively large, then the natural convective currents will be significant and the heat transfer coefficient will be high. Assisted circulation (such as stirring) that will induce forced convection, will also result in

higher coefficients. As convection is partially dependent on the bulk motion of the fluid, the viscosity (which varies with temperature) also has an important bearing on the thermal boundary layer. Additional variations can also occur on the steam side of the coil, especially with long lengths of pipe. The coil inlet may have a high steam velocity and may be relatively free from water. However, further along the length of the coil the steam velocity may be lower, and the coil may be running partially full of water. In very long coils, such as those sometimes found in seagoing tankers or in large bulk storage tanks, a significant pressure drop occurs along the length of the coil. To acheive the mean coil temperature, an average steam pressure of approximately 75% of the inlet pressure may be used. In extreme cases the average pressure used may be as low as 40% of the inlet pressure. Another variable is the coil material itself. The thermal conductivity of the coil material may vary considerably. However, overall heat transfer is governed to a large extent by the heat resistant films, and the thermal conductivity of the coil material is not as significant as their combined effect. Table 2.10.1 provides typical overall heat transfer coefficients for various conditions of submerged steam coil application. 'U' values for steam pressures between 2 bar g and 6 bar g should be found by interpolation of the data in the table.

The range of figures shown in Table 2.10.1 demonstrates the difficulty in providing definitive 'U' values. Customary figures at the higher end of the scale will apply to installations that are supplied with clean dry steam, small coils and good condensate drainage. The lower end is more applicable to poor quality steam, long coils and poor condensate drainage. The recommended overall heat transfer coefficients will apply to typical conditions and installations. These recommended rates are empirically derived, and will generally ensure that a generous safety margin applies to the coil sizing. In the case of fluids other than water, the heat transfer coefficient will vary even more widely due to the way in which viscosity varies with temperature. However, the values shown in Table 2.10.2 will serve as a guide for some commonly encountered substances, while Table 2.10.3 gives typical surface areas of pipes per metre length.

Example 2.10.1 Continuing from Example 2.9.1 determine: Part 1. The average steam mass flowrate during start-up. (Mean heat load = 367 kW) Part 2. The heat transfer area required. Part 3. A recommended coil surface area. Part 4. The maximum steam mass flowrate with the recommended heat transfer area. Part 5. A recommendation for installation, including coil diameter and layout. The following additional information has been provided:

Steam pressure onto the control valve = 2.6 bar g (3.6 bar a). A stainless steel steam coil provides heat. Heat transfer coefficient from steam/coil/liquid, U = 650 W/mC

Part 1 Calculate the average steam mass flowrate during start-up Steam pressure onto the control valve = 2.6 bar g (3.6 bar a) Critical pressure drop (CPD) will occur across the control valve during start-up, therefore the

minimum steam pressure in the heating coil should be taken as 58% of upstream absolute pressure. An explanation of this is given in Block 5.

Part 2 Calculate the heat transfer area required.

Part 3 A recommendation for coil surface area Because of the difficulties in providing accurate 'U' values, and to allow for future fouling of the heat exchange surface, it is usual to add 10% to the calculated heat transfer area.

Part 4 The maximum steam mass flowrate with the recommended heat transfer area Maximum heat transfer (and hence steam demand) will occur when the temperature difference between the steam and the process fluid is at its maximum, and should take into consideration the extra pipe area allowed for fouling. (a) Consider the maximum heating capacity of the coil Using Equation 2.5.3: = UAT
(coil)

(b) Steam flowrate to deliver 519 kW

Part 5 A recommendation for installation, including coil diameter and layout (a) Determine coil diameter and length

From Table 2.10.3, a 100 mm pipe has a surface area of 0.358 m/m run. This application will require:

It may be difficult to accommodate this length of large bore heating pipe to install in a 3 m 3 m tank. One solution would be to run a bank of parallel pipes between steam and condensate manifolds, set at different heights to encourage condensate to run to the lower (condensate) manifold. The drain line must fall from the bottom of the condensate manifold down to the steam trap (or pump-trap). See Figure 2.10.1 for a suggested layout.

Fig. 2.10.1 Possible layout of coils in a rectangular tank Note the steam supply is situated at one end of its manifold, whilst the trap set is at the other end. This will help steam to flow and push condensate through the coils. In the application, the steam and condensate headers would each be 2.8 m long. As the condensate manifold is holding condensate, the heat from it will be small compared to the steam manifold and this can be ignored in the calculation. The steam manifold should be 100 mm diameter as determined by the previous velocity calculation. This will provide a heating area of: 2.8 m x 0.358 m/m = 1.0 m Consequently 7 m - 1 m = 6 m of heat transfer area is still required, and must be provided by the connecting pipes. Arbitrarily selecting 32 mm pipe as a good compromise between robustness and workability:

The lengths of the connecting pipes are 2.5 m.

CHECK It is necessary to confirm the steam velocity through the connecting tubes: On the basis of proportionality of heat transfer area, the steam header will condense:

This leaves 86% of the 850 kg/h = 731 kg/h of steam which must pass through the 18 connecting pipes and also into the lower (condensate) manifold.

Other steam coil layouts The design and layout of the steam coil will depend on the process fluid being heated. When the process fluid to be heated is a corrosive solution, it is normally recommended that the coil inlet and outlet connections are taken over the lip of the tank, as it is not normally advisable to drill through the corrosion resistant linings of the tank side. This will ensure that there are no weak points in the tank lining, where there is a risk of leakage of corrosive liquids. In these cases the coil itself may also be made of corrosion resistant material such as lead covered steel or copper, or alloys such as titanium. However, where there is no danger of corrosion, lifts over the tank structure should be avoided, and the steam inlet and outlet connections may be taken through the tank side. The presence of any lift will result in waterlogging of a proportion of the coil length, and possibly waterhammer, noise and leaking pipework. Steam heating coils should generally have a gradual fall from the inlet to the outlet to ensure that condensate runs toward the outlet and does not collect in the bottom of the coil. Where a lift is unavoidable, it should be designed to include a seal arrangement at the bottom of the lift and a small bore dip pipe, as shown in Figure 2.10.2.

Fig. 2.10.2 Tank with a rising discharge pipe The seal arrangement allows a small amount of condensate to collect to act as a water seal, and prevents the occurrence of steam locking. Without this seal, steam can pass over any condensate collecting in the bottom of the pipe, and close the steam trap at the top of the riser. The condensate level would then rise and form a temporary water seal, locking the steam between the bottom of the riser and the steam trap. The steam trap remains closed until the locked steam condenses, during which time the coil continues to waterlog. When the locked steam condenses and the steam trap opens, a slug of water is discharged up the riser. As soon as the water seal is broken, steam will enter the rising pipe and close the trap, while the broken column of water falls back to lie at the bottom of the heating coil. The small bore dip pipe will only allow a very small volume of steam to become locked in the riser. It enables the water column to be easily maintained without steam bubbling through it, ensuring there is a steady and continuous condensate flow to the outlet. When the seal is ultimately broken, a smaller volume of water will return to the heating coil than with an unrestricted large bore riser, but as the water seal arrangement requires a smaller volume of condensate to form a water seal, it will immediately re-form. If the process involves articles being dipped into the liquid, it may not be convenient to install the coil at the bottom of the tank - it may be damaged by the objects being immersed in the solution. Also, during certain processes, heavy deposits will settle at the bottom of the tank and can quickly cover the heating surface, inhibiting heat transfer. For these reasons side hung coils are often used in the electroplating industry. In such cases serpentine or plate-type coils are arranged down the side of a tank, as shown in Figure 2.10.3.

These coils should also have a fall to the bottom with a water seal and a small bore dip-pipe. This arrangement has the advantage that it is often easier to install, and also easier to remove for periodic cleaning if required.

Fig. 2.10.3 Side hung coils If articles are to be dipped into the tank, it may not be possible to use any sort of agitator to induce forced convection and prevent temperature gradients occurring throughout the tank. Whether bottom or side coils are used, it is essential that they are arranged with adequate coverage so that the heat is distributed evenly throughout the bulk of the liquid. The diameter of the coil should provide sufficient length of coil for good distribution. A short length of coil with a large diameter may not provide adequate temperature distribution. However a very long continuous length of coil may experience a temperature gradient due to the pressure drop from end to end, resulting in uneven heating of the liquid. Whilst the next two headings, 'Sizing the control valve' and 'The condensate removal device' are included in this Tutorial, the new reader should refer to later Tutorials for full and comprehensive information, before attempting sizing and selection of equipment.

Control valve arrangement


The control valve set may be either one or two valves in parallel. A single control valve, large enough to cope with the maximum flowrate encountered at start-up, may be unable to control flow accurately at the minimum expected flowrate. This could cause erratic temperature control. An alternative is to fit two temperature control valves in parallel:

One valve (running valve) sized to control at the lower flowrate. A second valve (starting valve) to pass the difference between the capacity of the first valve, and the maximum flowrate.

The starting valve would have a set-point slightly lower than the running valve, so it would close first, leaving the running valve to control at low loads.

Sizing the control valve


The control valve set (either one valve or two valves in parallel). The coil has been sized on mean heat transfer values. However, it may be better to size the control valve to supply the maximum (start-up) load. With large coils in tanks, this will help to maintain a degree of steam pressure throughout the length of the coil when the steam is turned on, helping to push condensate through the coil to the steam trapping device. If the control valve were sized on mean values, steam pressure in the coil at start-up will tend to be lower and the coil may flood. Using one valve Continuing with Example 2.10.1 the maximum steam load is 850 kg/h and the coil is designed to deliver this at a pressure of 1.1 bar g. A steam valve sizing chart would show that a Kv of about 20 is required to pass 850 kg/h of steam with a pressure of 2.6 bar g at the inlet of the control valve, and Critical Pressure Drop (CPD) across the valve. (Tutorial 6.4 will show how the valve size can be determined by calculation). A DN40 control valve with a larger Kvs of 25 would therefore need to be selected for the application. If one valve is to be used, this valve must ensure the maximum heat load is catered for, while maintaining the required steam pressure in the coil to assist the drainage of condensate from it at start-up. However, for reasons previously explained, two valves may be better. The running load is 52 kW and with the coil running at 1.1 bar g, the running steam load:

The steam valve sizing chart shows a Kv of 2 is required to pass 85 kg/h with 3.6 bar upstream, operating at critical pressure drop. A DN15 KE type valve (Kvs = 4) and a DN25 piston actuated valve (Kvs = 18.6) operating together will cater for the start-up load. When approaching the control temperature, the larger valve would be set to shut down, allowing the smaller valve to give good control.

The condensate removal device

The selection and sizing of the condensate removal device will be very much influenced by the condensate backpressure. For the purpose of this example, it is assumed the backpressure is atmospheric pressure. The device should be sized so it is able to satisfy both of the following conditions:

Pass 850 kg/h of condensate with 1.1 bar g in the coil, i.e. the full-load condition. Pass the condensate load when steam pressure in the coil equals the condensate backpressure, i.e. the stall load condition.

If the steam trap is only sized on the first condition, it is possible that it may not pass the stall load (the condition where the product approaches its required temperature and the control valve modulates to reduce steam pressure). The stall load may be considerable. With respect to nonflow type applications such as tanks, this may not be too serious from a thermal viewpoint because the contents of the tank will almost be at the required temperature, and have a huge reservoir of heat. Any reduction in heat transfer at this part of the heating process may therefore have little immediate effect on the tank contents. However, condensate will back up into the coil and waterhammer will occur, along with its associated symptoms and mechanical stresses. Tank coils in large circular tanks tend to be of robust construction, and are often able to withstand such stresses. Problems can however occur in rectangular tanks (which tend to be smaller), where vibration in the coil will have more of an effect on the tank structure. Here, the energy dissipated by the waterhammer causes vibration, which can be detrimental to the life of the coil, the tank, and the steam trap, as well as creating unpleasant noise. With respect to flow-type applications such as plate heat exchangers, a failure to consider the stall condition will usually have serious implications. This is mainly due to the small volume in the heat exchanger. For heat exchangers, any unwanted reduction in the heating surface area, such as that caused by condensate backing up into the steam space, can affect the flow of heat through the heating surface. This can cause the control system to become erratic and unstable, and processes requiring stable or accurate control can suffer with poor performance. If heat exchangers are oversized, sufficient heating surface may remain when condensate backs up into the steam space, and reduction of thermal performance may not always occur. However, with heat exchangers not designed to cope with the effects of waterlogging, this can lead to corrosion of the heating surface, inevitably reducing the service life of the exchanger. Waterlogging can, in some applications, be costly. Consider a waterlogging air heater frost coil. Cold air at 4C flowing at 3 m/s can soon freeze condensate locked in the coils, resulting in premature and unwarranted failure. Proper drainage of condensate is essential to maintain the service life of any heat exchanger and air heater. Steam traps are devices which modulate to allow varying amounts of condensate to drain from

applications under varying conditions. Float traps are steam traps designed to modulate and release condensate close to steam temperature, offering maximum plant performance, maximum plant life, and maximum return on plant investment. When stall conditions occur, and a steam trap cannot be used, an automatic pump-trap or pump and trap in combination will ensure correct condensate drainage at all times, thus maximising the thermal capability and lifetime costs of the plant.

Steam jackets
The most commonly used type of steam jacket consists simply of an outer cylinder surrounding the vessel, as shown in Figure 2.10.4. Steam circulates in the outer jacket, and condenses on the wall of the vessel. Jacketed vessels may also be lagged, or may contain an internal air space surrounding the jacket. This is to ensure that as little steam as possible condenses on the outer jacket wall, and that the heat is transferred inwards to the vessel.

Fig. 2.10.4 A conventional jacketed vessel The heat transfer area (the vessel wall surface area), can be calculated in the same manner as with a steam coil, using Equation 2.5.3 and the overall heat transfer coefficients provided in Table 2.10.4. Although steam jackets may generally be less thermally efficient than submerged coils, due to radiation losses to the surroundings, they do allow space for the vessels to be agitated so that heat transfer is promoted. The U values listed in Table 2.10.4. are for moderate non-proximity agitation. Commonly the vessel walls are made from stainless steel or glass lined carbon steel. The glass lining will offer an additional corrosion resistant layer. The size of the steam jacket space will depend on the size of the vessel, but typically the width may be between 50 mm and 300 mm.

Potrebbero piacerti anche