Sie sind auf Seite 1von 326

A Tutorial on Power Generation From Thermal Power Plants - TS Graph, Process, Principles, Turbine Model, Basic Boiler Schematics

Power Generation from Thermal Power Plant Introduction About 70 % of energy used by India is produced in Coal fired thermal power plants. Not just India, people all over the world heavily rely on thermal power stations. This is because of the abundant availability of coal, reliable cheap power and early advent of steam engine technology. Though there is a lot of hue and cry over the CO2 emissions and diminishing coal reserves, coal power continues to dominate the energy sector. Rankine cycle is the working principle of the plants all over the world. Water is boiled into steam which is super heated. This is the phase where the energy of the coal is give to the steam/water. The high pressure and high temperature steam is allowed to expand in turbines coupled with generators. Here , a part of energy is given back by the steam. Most of remaining heat is dissipated to atmosphere. More about this will be discussed in the efficiency discussion.

The TS graph for Basic Rankine cycle For an ideal plant, there are good number of specifications to be satisfied. The power plant must be located to a coal mine as close as possible. If the plant is dependent on the imported coal it should be closely located to the sea port. In either cases, dedicated transporation system must exist for transmission if coal reserves. Another important aspect is the ash disposable facility. Indian Coal has a higgm amount of ash content which turns out to be around 30 -40 %. This if not disposed properly, results in health hazards in and around the plant leading to numerous other problems. Presently the ash is used for various industries and also used for domestic purposes. In most cases it is stored in propoer places. Huge quantities of water must be required for condenser, disposal of ash and feed water circuit etc. It is therefore desirable to locate plant on side of river. For example, VTPS in Vijayawada is located on the banks of river Krishna. The Process and Principle

The basic entities in the power plant are Boiler, Turbine , Condenser , pump. Here, the principle is explained with the help of Temperature and Entropy (TS) curve. Starting at 1, the water at room temperature is boiled at constant temperature in the boiler. This process has a T constant for it is the boiling of water which takes at a constant temperature. Here we are increasing the entropy of by phase change and the energy is sorted in the form of latent heat also. As you might assume, the temperature is not 1000c because this boiler is at higher pressure and the boiling point is also high. As shown in the figure the temperature is constant and the line is parallel to S axis. All this process happens in a boiler. The end product is steam.

TS graph for the power plant operation Once you have steam, you cannot immediately pump it to the turbine to extract energy because once it enters turbine and starts condensing forming water droplets without giving much energy. The point here is, just a phase change wont work out for energy tr ansfer from coal in boiler to turbine. The steam must be heated to higher temperatures. This phase is called super heater and this phase is mainly executed in super heater. This super heater will be located in between boiler and the turbine. The source of heat for the super heater is the hot flue gases obtained in the boiler after burning coal. This explains the phase 2-3 in the T-S curve. Note that , in this process, neither the temperature nor the entropy remains constant. Now, in the next process, the super heated steam is allowed to expand in the turbine. As, the high pressure steam is allowed through a small nozzle ,steam acquires kinetic energy. This kinetic energy of the steam will exert required force on the turbine blades. The turbine is designed well to receive maximum force from the steam. This process is a constant entropy process. The steam is allowed inside the turbine until water droplets begin to form. In practice, formation of water drops is strictly prohibited for the water drops will impinge the turbine blades and cause corrosion. Now the ouput of the turbine is low pressure and low temperature steam. This accounts for the phase 3-4 of the cycle. In this phase of condenser, the heat of steam is exchanged with a heat exchanger, essentially water. The steam now turns into water and this is processed again and sent into boiler for the next

cycle. The heat exchanger gets heated up and this needs to be cooled for further use. Hence the heat of exchanger is dissipated in atmosphere through large cooling towers. A lot of energy from the entire system remains unused in this 4-5 . The analysis can be obtained in the efficiency discussion at the end of the tutorial. As you have seen throughout the process, water needs to be flown from one part to other. The necessary draught is created by the pump. The step 5-6 is a pump which is used to circulate water. During this process, a little temperature change can be observed. Finally, the cooled water cannot be directly sent into boiler. Because the boiler is at higher temperature, it causes irregularly expansion which results in collapse of the boiler. Hence, the water should be heated to higher temperature. This is done in economizer which uses heat from flue gases. Thus this accounts for the 6-1 phase of the Rankine cycle. The efficiency of the Rankine cycle is given by 1-T2T1. Where T2 is the temperature of super heated steam and the T1 is the temperature of the water entering inside. Basic Flow Chart of Power Plant Construction of the plant and parts. As cited above, the primary parts of the plant are the boiler, turbine, condenser are explained below. Other numerous parts including pulveriser, water treatment plant, cooling towers etc are also discussed in detail. Boiler : It is used to convert the water into steam where coal is burnt. It is a relatively huge structure with a typical boiler of a 500 MW plant would be equivalent to 5 storied building. The boiler material will mostly made of cast iron to with stand high temperature and pressure. The construction of boiler varies depending upon the heat transfer method used. In a traditional boiler, the boiler has holes on the lower bottom for the coil powder to enter. The coal enters in such a way that , it creates a vortex inside the boiler. This is to ensure that coal spends maximum time before settling down and gets burnt completely. The outer surface of the boiler has thousands of pipes in which water runs through. This is the process in which heat is exchanged. The flue gases rising out of burning coal pass through the super heater as shown in the figure.

Basic Boiler Schematic

Turbine : Steam turbines convert the energy acquired by the steam in to the mechanical energy. These turbines are couple with the alternators which produce the electrical energy from the mechanical energy. Two types of turbines are widely prevalent : impulse turbines and reaction turbines. In the impulse turbines, steam expands at the nozzles and achieved kinetic energy is used to rotate the blades of the turbine. The blades change the direction of steam but not the pressure. Thus change in momentum can be accounted for rotation of the rotor. In the reaction turbines, steam is partially expanded on the nozzles and remaining expansion takes place during the flow over moving blades. Generally there are two or three sets of turbines at one go. All the enrgy stored in the steam cannot be obtained at one go from a single turbine. So, there are two or three sets of turbines located which are connected by a shaft. Now the high pressure steam enters into the first turbine, lets call it HP turbine. Once the expansion takes place, the pressure falls. Hence we need to use a turbine designed for lower pressure appropriate to the out coming steam. So the second turbine will be a medium pressure turbine (MP turbine). Further , in some cases a third turbine is also added to make more energy out of steam and this is called a low pressure turbine. The specifications of turbines are calclcuated during the plant design and later during operations, same ratings of steam pressure and temperature need to be maintained for optimum operation. Given below is a figure illustrating the construction of the three

stage turbine. Three turbine model for a plant Pulverizer : To generate the massive amount of heat which is required instantaneously , a lot of coal is to be burnt. If chunks of coal is used, very less surface area of coal is exposed and it takes a lot of space to burn enough coal chunks for required power. As a result, to overcome this problem, coal is pulverized into powder which is as smooth as talc. Now this powder is blown into the boiler. Thus, as powder has a very higher surface area compared to chunks of coal, very large amount of coal can be burnt instantaneously in less volume efficiently. This is the underlying interesting fact for using of a pulveriser.

Super heater : The steam is super heated in order to make it hold more energy and transfer it to the turbine. This job is accomplished by the super heater. Super heater is showed in the boiler schematic. The flue gases coming out of the boiler are used to super heat the steam.

Economizer : The water entering into the boiler must have a temperature compatible with the boiler temperature. So, the heat left with the flue gases after super heater is used to heat the water in the economizer. The economizer has convoluted tubes in which water flows and the flue gases flow over these tubes in a closed structure. Air pre heater : The air used for combustion of the coal is also pre heated by the flue gases so as to take maximum heat from the gases before they diffuse in to the atmosphere. It is also to ensure that the un heated air should not interfere with proper combustion inside the boiler.Re heater : To improve the efficiency of the plant, there is something interesting done. The area under curve is the output or work done. If we could improve the area, we can improve the efficiency of the system. The steam which comes out of high pressure turbine is taken out and heated using flue gases and this reheated steam is sent into IP turbine. As a result the new TS graph looks like below.

Improved TS graph after using re heater Condenser : As discussed earlier, the job of condenser is to turn the steam from the turbine into water and thereby reducing the amount of water required for each cycle. There are many types of condensers. The familiar ones are Jet type and Surface type. In the jet type , the cooling water and the steam are mixed and the resultant steam water mixture is drawn outside. Surface type uses a different circuit for both and the steam is converted into water and cooling water turns hot. The surface type are the widely prevalent ones.

Generators : The generators also called alternators are coupled with turbines which generate electrical energy. The output of the generator at 11KV is stepped up to higher voltage of 220KV and transmitted through the transmission lines. Here , the interesting area of study is to control the output power of the generator. As the load on the system continuously vary and as the energy cannot be stored, the output of generator has to be varied according to load. This aspect will be covered in Power Systems Operations and Control tutorial.

Overview of a practical power plant Miscellaneous parts :

Ash handling plant, Ash precipitators, pumps for draught, turbine governing system etc. The Efficiency discussion The efficiency of the thermal powerplant calculated from the Rankine cycle will be around 45% . But practically efficiencies of 35%-38% only have been achieved so far. This is due to various losses present in the entire system when put into practice. Out of the losses, energy lost as heat takes a major chunk. This energy is lost at mainly two points: flue gases entering into atmosphere and cooling of the condensate. The cooling of the condensate is a part and parcel of the cycle and nothing can be done there to increase the efficiency. Now we can consider decreasing the temperature of flue gases as much as possible up to room temperatures. But it is also disadvantageous , for, the cooled gases do not flow outside the boiler circuit on their own. They need forced draught to go out if the temperature is equal to room temperature. So there is a lower limit for temperature of flue gases. Also note that, high pressure cannot be maintained inside the boiler. It is recommended to maintain slightly low pressure in the boiler otherwise, the fire will come out from every possible gap in the boiler. Pollution aspects : Thermal energy is the most unclean energy. It causes thermal pollution and air pollution apart from leaving off a lot of ash. The ash can be used for other purposes or should be disposed properly otherwise during dry season , it mixes with air and makes the surrounding places uncomfortable to live. The plant also produces thermal pollution ie by adding more and more heat to the atmosphere. But as Nature is a huge sink of heat this doesnt add much trouble. Other pollution from the plant is due to production of soot, SO x , COx gases and consequent problems. Nowadays, latest technologies are being implemented to minimize the emission of these gases by designing the boiler in a special way and adding other compounds so as to neutralize these gases. Problem : Find out the theoretical efficiency of a power plant whose steam is heated up to a temperature of 4000 Celsius and water temperature at the initial stages is 75 0 Celsius. Efficiency : 1-(75+273)/(400+273) = 1- 0.51 = 0.49 = 49 % efficiency.

8.6 Enhancements of, and Effect of Design Parameters on, Rankine Cycles
The basic Rankine cycle can be enhanced through processes such as superheating and reheat. Diagrams for a Rankine cycle with superheating are given in Figure 8.13. The heat addition is continued past the point of vapor saturation, in other words the vapor is heated so that its temperature is higher than the saturation temperature associated with . This does several things. First, it increases the mean temperature at which heat is added, , thus increasing the efficiency of the cycle. Second is that the quality of the two-phase mixture during the expansion is higher with superheating, so that there is less moisture content in the mixture as it flows through the turbine. (The moisture content at is less than that at .) This is an advantage in terms of decreasing the mechanical deterioration of the blading.

coordinates]

coordinates]

coordinates]

Figure 8.13: Rankine cycle with superheating

The heat exchanges in the superheated cycle are:


Along Along :

, which is a constant pressure (isobaric) process: , .

The thermal efficiency of the ideal Rankine cycle with superheating is

This can be expressed explicitly in terms of turbine work and compression (pump) work as

Compared to the basic cycle, superheating has increased the turbine work, increased the mean temperature at which heat is received, the cycle efficiency. , and increased

Figure 8.14: Comparison of Rankine cycle with superheating and Carnot cycle

Figure 8.15: Rankine cycle with superheating and reheat for space power application

A comparison of the Carnot cycle and the Rankine cycle with superheating is given in Figure 8.14. The maximum and minimum temperatures are the same, but the average temperature at which heat is absorbed is lower for the Rankine cycle. To alleviate the problem of having moisture in the turbine, one can heat again after an initial expansion in a turbine, as shown in Figure 8.15, which gives a schematic of a Rankine cycle for space power application. This process is known as reheat. The main practical advantage of reheat (and of superheating) is the decrease in moisture content in the turbine because most of the heat addition in the cycle occurs in the vaporization part of the heat addition process.

Figure 8.16: Effect of exit pressure on Rankine cycle efficiency

We can also examine the effect of variations in design parameters on the Rankine cycle. Consider first the changes in cycle output due to a decrease in exit pressure. In terms of the cycle shown in Figure 8.16, the exit pressure would be decreased from to . The original cycle is , and the

modified cycle is . The consequences are that the cycle work, which is the integral of around the cycle, is increased. In addition, as drawn, although the levels of the mean temperature at which the heat is absorbed and rejected both decrease, the largest change is the mean temperature of the heat rejection, so that the thermal efficiency increases.

Figure 8.17: Effect of maximum boiler pressure on Rankine cycle efficiency

Another design parameter is the maximum cycle pressure. Figure 8.17 shows a comparison of two cycles with different maximum pressure but the same maximum temperature, which is set by material properties. The average temperature at which the heat is supplied for the cycle with a higher maximum pressure is increased over the original cycle, so that the efficiency increases. Muddy Points Why do we look at the ratio of pump (compression) work to turbine work? We did not do that for the Brayton cycle. (MP 8.10)

Shouldn't the efficiency of the super/re-heated Rankine cycle be larger because its area is greater? (MP 8.11) Why can't we harness the energy in the warm water after condensing the steam in a power plant? (MP 8.12)

5.1 Concept and Statements of the Second Law (Why do we need a second law?)
The unrestrained expansion, or the temperature equilibration of the two bricks, are familiar processes. Suppose you are asked whether you have ever seen the reverse of these processes take place? Do two bricks at a medium temperature ever go to a state where one is hot and one is cold? Will the gas in the unrestrained expansion ever spontaneously return to occupying only the left side of the volume? Experience hints that the answer is no. However, both these processes, unfamiliar though they may be, are compatible with the first law. In other words the first law does not prohibit their occurrence. There thus must be some other ``great principle'' that describes the direction of natural processes, that tells us which first law compatible processes will not be observed. This is contained in the second law. Like the first law, it is a generalization from an enormous amount of observation. There are several ways in which the second law of thermodynamics can be stated. Listed below are three that are often encountered. As described in class (and as derived in almost every thermodynamics textbook), although the three may not appear to have much connection with each other, they are equivalent. 1. No process is possible whose sole result is the absorption of heat from a reservoir and the conversion of this heat into work. [Kelvin-Planck statement of the second law]

Figure 5.1: This is not possible (Kelvin-Planck)

2. No process is possible whose sole result is the transfer of heat from a cooler to a hotter body. [Clausius statement of the second law]

Figure 5.2: For

, this is not possible (Clausius)

3. There exists for every system in equilibrium a property called entropy, , which is a thermodynamic property of a system. For a reversible process, changes in this property are given by

The entropy change of any system and its surroundings, considered together, is positive and approaches zero for any process which approaches reversibility.

For an isolated system, i.e., a system that has no interaction with the surroundings, changes in the system have no effect on the surroundings. In this case, we need to consider the system only, and the first and second laws become:

For an isolated system the total energy ( ) is constant. The entropy can only increase or, in the limit of a reversible process, remain constant.

The limit, or , represents the best that can be done. In thermodynamics, propulsion, and power generation systems we often compare performance to this limit to measure how close to ideal a given process is. All of these statements are equivalent, but 3 gives a direct, quantitative measure of the departure from reversibility. Entropy is not a familiar concept and it may be helpful to provide some additional rationale for its appearance. If we look at the first law,

the term on the left is a function of state, while the two terms on the right are not. For a simple compressible substance, however, we can write the work done in a reversible process as , so that

Two out of the three terms in this equation are expressed in terms of state variables. It seems plausible that we ought to be able to express the third term using state variables as well, but what are the appropriate variables? If so, the term should perhaps be viewed as analogous to where the parentheses denote an intensive state variable and the square brackets denote an extensive state variable. The second law tells us that the intensive variable is the temperature, , and the extensive state variable is the entropy, . The first law for a simple compressible substance in terms of state variables is thus
(5..1)

Because Eq. 5.1 includes the second law, it is referred to as the combined first and second law. Because it is written in terms of state variables, it is true for all processes, not just reversible ones. We summarize below some attributes of entropy:

1. Entropy is a function of the state of the system and can be found if any two properties of the system are known, e.g. or or

. 2. is an extensive variable. The entropy per unit mass, or specific entropy, is . 3. The units of entropy are Joules per degree Kelvin (J/K). The units for specific entropy are J/K-kg. 4. For a system, , where the numerator is the heat given to the system and the denominator is the temperature of the system at the location where the heat is received. 5. for pure work transfer.

Muddy Points Why is What makes always true? (MP 5.1) different than ? (MP 5.2)

Thermodynamics and Propulsion

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements of Contents Index

5.3 Combined First and Second Law Expressions


The first law, written in a form that is always true:

For reversible processes only, work or heat may be rewritten as

Substitution leads to other forms of the first law true for reversible processes only:

(If the substance has other work modes, e.g., stress, strain,

where

is a pressure-like quantity, and

is a volume-like quantity.)

Substituting for both

and

in terms of state variables,

The above is always true because it is a relation between properties and is now independent of process.

In terms of specific quantities:

The combined first and second law expressions are often more usefully written in terms of the enthalpy, or specific enthalpy, ,

Or, since

In terms of enthalpy (rather than specific enthalpy) the relation is

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements of Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel Cycle Contents Index Subsections

3.7.1 Work and Efficiency 3.7.2 Gas Turbine Technology and Thermodynamics 3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet 3.7.4 MIT Cogenerator

3.7 Brayton Cycle


[VW, S & B: 9.8-9.9, 9.12]

The Brayton cycle (or Joule cycle) represents the operation of a gas turbine engine. The cycle consists of four processes, as shown in Figure 3.13 alongside a sketch of an engine:

a - b Adiabatic, quasi-static (or reversible) compression in the inlet and compressor; b - c Constant pressure fuel combustion (idealized as constant pressure heat addition); c - d Adiabatic, quasi-static (or reversible) expansion in the turbine and exhaust nozzle, with which we 1. take some work out of the air and use it to drive the compressor, and

2. take the remaining work out and use it to accelerate fluid for jet propulsion, or to turn a generator for electrical power generation; d - a Cool the air at constant pressure back to its initial condition.

Figure 3.13: Sketch of the jet engine components and corresponding thermodynamic states

The components of a Brayton cycle device for jet propulsion are shown in Figure 3.14. We will typically represent these components schematically, as in Figure 3.15. In practice, real Brayton cycles take one of two forms. Figure 3.16(a) shows an ``open'' cycle, where the working fluid enters and then exits the device. This is the way a jet propulsion cycle works. Figure 3.16(b) shows the alternative, a closed cycle, which recirculates the working fluid. Closed cycles are used, for example, in space power generation.

Figure 3.14: Schematics of typical military gas turbine engines. Top: turbojet with afterburning, bottom: GE F404 low bypass ratio turbofan with afterburning (Hill and Peterson, 1992).

Figure 3.15: Thermodynamic model of gas turbine engine cycle for power generation

[Open cycle operation]

[Closed cycle

operation] Figure 3.16: Options for operating Brayton cycle gas turbine engines

Muddy Points Would it be practical to run a Brayton cycle in reverse and use it as refrigerator? (MP 3.10)

3.7.1 Work and Efficiency


The objective now is to find the work done, the heat absorbed, and the thermal efficiency of

the cycle. Tracing the path shown around the cycle from first law gives (writing the equation in terms of a unit mass),

and back to

, the

Here is zero because is a function of state, and any cycle returns the system to its starting 3.2 state . The net work done is therefore

where , are defined as heat received by the system ( evaluate the heat transferred in processes - and - .

is negative). We thus need to

For a constant pressure, quasi-static process the heat exchange per unit mass is

We can see this by writing the first law in terms of enthalpy (see Section 2.3.4) or by remembering the definition of .

The heat exchange can be expressed in terms of enthalpy differences between the relevant states. Treating the working fluid as a perfect gas with constant specific heats, for the heat addition from the combustor,

The heat rejected is, similarly,

The net work per unit mass is given by

The thermal efficiency of the Brayton cycle can now be expressed in terms of the temperatures:
(3..8 )

To proceed further, we need to examine the relationships between the different

temperatures. We know that points points and , and and reversible, so ;

and

are on a constant pressure process as are . The other two legs of the cycle are adiabatic

Therefore , or, finally, . Using this relation in the expression for thermal efficiency, Eq. (3.8) yields an expression for the thermal efficiency of a Brayton cycle:

(3..9)

The temperature ratio across the compressor, . In terms of compressor temperature ratio, and using the relation for an adiabatic reversible process we can write the efficiency in terms of the compressor (and cycle) pressure ratio, which is the parameter commonly used:

(3..10)

Figure 3.17: Gas turbine engine pressures and temperatures

Figure 3.17 shows pressures and temperatures through a gas turbine engine (the PW4000, which powers the 747 and the 767).

Figure 3.18: Gas turbine engine pressure ratio trends (Janes Aeroengines, 1998)

Figure 3.19: Trend of Brayton cycle thermal efficiency with compressor pressure ratio

Equation (3.10) says that for a high cycle efficiency, the pressure ratio of the cycle should be increased. This trend is plotted in Figure 3.19. Figure 3.18 shows the history of aircraft engine pressure ratio versus entry into service, and it can be seen that there has been a large increase in cycle pressure ratio. The thermodynamic concepts apply to the behavior of real aerospace devices!

Muddy Points When flow is accelerated in a nozzle, doesn't that reduce the internal energy of the flow and therefore the enthalpy? (MP 3.11) Why do we say the combustion in a gas turbine engine is constant pressure? (MP 3.12) Why is the Brayton cycle less efficient than the Carnot cycle? (MP 3.13) If the gas undergoes constant pressure cooling in the exhaust outside the engine, is that still within the system boundary? (MP 3.14) Does it matter what labels we put on the corners of the cycle or not? (MP 3.15) Is the work done in the compressor always equal to the work done in the turbine plus work out (for a Brayton cyle)? (MP 3.16)

3.7.2 Gas Turbine Technology and Thermodynamics


The turbine entry temperature, , is fixed by materials technology and cost. (If the temperature is too high, the blades fail.) Figures 3.20 and 3.21 show the progression of the turbine entry temperatures in aeroengines. Figure 3.20 is from Rolls Royce and Figure 3.21 is from Pratt & Whitney. Note the relation between the gas temperature coming into the turbine blades and the blade melting temperature.

Figure 3.20: Rolls-Royce high temperature technology

Figure 3.21: Turbine blade cooling technology [Pratt & Whitney]

For a given level of turbine technology (in other words given maximum temperature) a design question is what should the compressor be? What criterion should be used to decide this? Maximum thermal efficiency? Maximum work? We examine this issue below.

Figure 3.22: Efficiency and work of two Brayton cycle engines

The problem is posed in Figure 3.22, which shows two Brayton cycles. For maximum

efficiency we would like as high as possible. This means that the compressor exit temperature approaches the turbine entry temperature. The net work will be less than the heat received; as the heat received approaches zero and so does the net work.

The net work in the cycle can also be expressed as

, evaluated in traversing the

cycle. This is the area enclosed by the curves, which is seen to approach zero as . The conclusion from either of these arguments is that a cycle designed for maximum thermal efficiency is not very useful in that the work (power) we get out of it is zero. A more useful criterion is that of maximum work per unit mass (maximum power per unit mass flow). This leads to compact propulsion devices. The work per unit mass is given by:

where

is the maximum turbine inlet temperature (a design constraint) and

is atmospheric

temperature. The design variable is the compressor exit temperature, as this is varied, we differentiate the expression for work with respect to

, and to find the maximum :

The first and the fourth terms on the right hand side of the above equation are both zero (the turbine entry temperature is fixed, as is the atmospheric temperature). The maximum work occurs where the derivative of work with respect to is zero:

(3..11)

To use Eq. (3.11), we need to relate

and

. We know that

Hence,

Plugging this expression for the derivative into Eq. (3.11) gives the compressor exit temperature for maximum work as . In terms of temperature ratio,

The condition for maximum work in a Brayton cycle is different than that for maximum efficiency. The role of the temperature ratio can be seen if we examine the work per unit mass which is delivered at this condition:

Ratioing all temperatures to the engine inlet temperature,

To find the power the engine can produce, we need to multiply the work per unit mass by the mass flow rate:
(3..12 )

The trend of work output vs. compressor pressure ratio, for different temperature ratios , is shown in Figure 3.23.

Figure 3.23: Trend of cycle work with compressor pressure ratio, for different temperature ratios

[Gas turbine engine core]

[Core power vs. turbine entry

temperature] Figure 3.24: Aeroengine core power [Koff/Meese, 1995]

Figure 3.24 shows the expression for power of an ideal cycle compared with data from

actual jet engines. Figure 3.24(a) shows the gas turbine engine layout including the core (compressor, burner, and turbine). Figure 3.24(b) shows the core power for a number of different engines as a function of the turbine rotor entry temperature. The equation in the figure for horsepower (HP) is the same as that which we just derived, except for the conversion factors. The analysis not only shows the qualitative trend very well but captures much of the quantitative behavior too. A final comment (for this section) on Brayton cycles concerns the value of the thermal efficiency. The Brayton cycle thermal efficiency contains the ratio of the compressor exit temperature to atmospheric temperature, so that the ratio is not based on the highest temperature in the cycle, as the Carnot efficiency is. For a given maximum cycle temperature, the Brayton cycle is therefore less efficient than a Carnot cycle.

Muddy Points

What are the units of

in

? (MP 3.17)

Question about the assumptions made in the Brayton cycle for maximum efficiency and maximum work (MP 3.18) You said that for a gas turbine engine modeled as a Brayton cycle the work done is , where is the heat added and is the heat rejected. Does this suggest that the work that you get out of the engine doesn't depend on how good your compressor and turbine are? since the compression and expansion were modeled as adiabatic. (MP 3.19)

3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet


A schematic of a ramjet is given in Figure 3.25.

Figure 3.25: Ideal ramjet [J. L. Kerrebrock, Aircraft Engines and Gas Turbines]

In the ramjet there are ``no moving parts.'' The processes that occur in this propulsion

device are:

: Isentropic diffusion (slowing down) and compression, with a decrease in Mach number, . : Constant pressure combustion. : Isentropic expansion through the nozzle.

The ramjet thermodynamic cycle efficiency can be written in terms of flight Mach number, , as follows:

and

so

See also Section 11.6.3 for other figures of merit.

Muddy Points Why don't we like the numbers 1 and 2 for the stations? Why do we go 0-3? (MP 3.20) For the Brayton cycle efficiency, why does ? (MP 3.21)

3.7.4 MIT Cogenerator


MIT operates a Brayton cycle power generator on campus. For more information, see the website at https://cogen.mit.edu/ctg.cfm .

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel Cycle Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2 Generalized Representation of Contents Index

3.3 The Carnot Cycle


A Carnot cycle is shown in Figure 3.4. It has four processes. There are two adiabatic reversible legs and two isothermal reversible legs. We can construct a Carnot cycle with many different systems, but the concepts can be shown using a familiar working fluid, the ideal gas. The system can be regarded as a chamber enclosed by a piston and filled with this ideal gas.

Figure 3.4: Carnot cycle -- thermodynamic diagram on left and schematic of the different stages in the cycle for a system composed of an ideal gas on the right

The four processes in the Carnot cycle are:


1. The system is at temperature at state . It is brought in contact with a heat reservoir, which is just a liquid or solid mass of large enough extent such that its temperature does not change appreciably when some amount of heat is transferred to the system. In other words, the heat reservoir is a constant temperature source (or receiver) of heat. The system then undergoes an isothermal expansion from to , with heat absorbed . 2. At state , the system is thermally insulated (removed from contact with the heat reservoir) and then let expand to . During this expansion the temperature decreases to ) . The

heat exchanged during this part of the cycle,

3. At state

the system is brought in contact with a heat reservoir at temperature

. It is

then compressed to state , rejecting heat in the process. 4. Finally, the system is compressed adiabatically back to the initial state exchange .

. The heat

The thermal efficiency of the cycle is given by the definition

(3..4)

In this equation, there is a sign convention implied. The quantities

as defined are

the magnitudes of the heat absorbed and rejected. The quantities , on the other hand are defined with reference to heat received by the system. In this example, the former is negative and the latter is positive. The heat absorbed and rejected by the system takes place during isothermal processes and we already know what their values are from Eq. (3.1):

The efficiency can now be written in terms of the volumes at the different states as (3..5)

The path from states to and from to reversible adiabatic process we know that of state, we have Along the curve ,

are both adiabatic and reversible. For a . Using the ideal gas equation , therefore, .

. Along curve . Thus,

Comparing the expression for thermal efficiency Eq. (3.4) with Eq. (3.5) shows two consequences. First, the heats received and rejected are related to the temperatures of the isothermal parts of the cycle by

(3..6)

Second, the efficiency of a Carnot cycle is given compactly by

(3..7)

The efficiency can be 100% only if the temperature at which the heat is rejected is zero. The heat and work transfers to and from the system are shown schematically in Figure 3.5.

Figure 3.5: Work and heat transfers in a Carnot cycle between two heat reservoirs

Muddy Points

Since the ,

, looking at the

graph, does that mean the farther apart

isotherms are, the greater efficiency? And that if they were very close, it would

be very inefficient? (MP 3.2) In the Carnot cycle, why are we only dealing with volume changes and not pressure changes on the adiabats and isotherms? (MP 3.3) Is there a physical application for the Carnot cycle? Can we design a Carnot engine for a propulsion device? (MP 3.4) How do we know which cycles to use as models for real processes? (MP 3.5)

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2 Generalized Representation of Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 8.8 Some Overall Comments Up: 8. Power Cycles with Previous: 8.6 Enhancements of Rankine Contents Index

8.7 Combined Cycles in Stationary Gas Turbine for Power Production


The turbine entry temperature in a gas turbine (Brayton) cycle is considerably higher than the peak steam temperature. Depending on the compression ratio of the gas turbine, the turbine exhaust temperature may be high enough to permit efficient generation of steam using the ``waste heat'' from the gas turbine. A configuration such as this is known as a gas turbine-steam combined cycle power plant. The cycle is illustrated in Figure 8.18.

Figure 8.18: Gas turbine-steam combined cycle [Kerrebrock,Aircraft Engines and Gas Turbines]

Figure 8.19: Schematic of combined cycle using gas turbine (Brayton cycle) and steam turbine (Rankine cycle) [Langston]

The heat input to the combined cycle is the same as that for the gas turbine, but the work output is larger (by the work of the Rankine cycle steam turbine). A schematic of the overall heat engine, which can be thought of as composed of an upper and a lower heat engine in series, is given in Figure 8.19. The upper engine is the gas turbine (Brayton cycle) which expels heat to the lower engine, the steam turbine (Rankine cycle). The overall efficiency of the combined cycle can be derived as follows. We denote the heat received by the gas turbine as and the heat rejected to the atmosphere as . The

heat out of the gas turbine is denoted as . The hot exhaust gases from the gas turbine pass through a heat exchanger where they are used as the heat source for the two-phase Rankine cycle, so that cycle efficiency is is also the heat input to the steam cycle. The overall combined

where the subscripts refer to combined cycle (CC), Brayton cycle (B) and Rankine cycle (R) respectively.

From the first law, the overall efficiency can be expressed in terms of the heat inputs and heat rejections of the two cycles as (using the quantity the heat transferred): to denote the magnitude of

The first square bracket term on the right hand side is the Brayton cycle efficiency, is the Rankine cycle efficiency, , and the term in parentheses is cycle efficiency can thus be written as

, the second

. The combined

(8..5)

Equation (8.5) gives insight into why combined cycles are so successful. Suppose that the gas turbine cycle has an efficiency of 40%, which is a representative value for current Brayton cycle gas turbines, and the Rankine cycle has an efficiency of 30%. The combined cycle efficiency would be 58%, which is a very large increase over either of the two simple cycles. Some representative efficiencies and power outputs for different cycles are shown in Figure 8.20.

Figure 8.20: Comparison of efficiency and power output of various power products [Bartol (1997)]

Next: 8.8 Some Overall Comments Up: 8. Power Cycles with Previous: 8.6 Enhancements of Rankine Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's

All Contents Index Subsections


1.2.1 The Continuum Model 1.2.2 The Concept of a ``System'' 1.2.3 The Concept of a ``State'' 1.2.4 The Concept of ``Equilibrium'' 1.2.5 The Concept of a ``Process'' 1.2.6 Quasi-Equilibrium Processes 1.2.7 Equations of state

1.2 Definitions and Fundamental Ideas of Thermodynamics


As with all sciences, thermodynamics is concerned with the mathematical modeling of the real world. In order that the mathematical deductions are consistent, we need some precise definitions of the basic concepts. The following is a discussion of some of the concepts we will need. Several of these will be further amplified in the lectures and in other handouts. If you need additional information or examples concerning these topics, they are described clearly and in-depth in (SB&VW). They are also covered, although in a less detailed manner, in Chapters 1 and 2 of the book by Van Ness.

1.2.1 The Continuum Model


Matter may be described at a molecular (or microscopic) level using the techniques of statistical mechanics and kinetic theory. For engineering purposes, however, we want ``averaged'' information, i.e., a macroscopic, not a microscopic, description. There are two reasons for this. First, a microscopic description of an engineering device may produce too much information to manage. For example, pressure contains of air at standard temperature and molecules (VW, S & B:2.2), each of which has a position and a

velocity. Typical engineering applications involve more than molecules. Second, and more importantly, microscopic positions and velocities are generally not useful for determining how macroscopic systems will act or react unless, for instance, their total effect is integrated. We therefore neglect the fact that real substances are composed of discrete molecules and model matter from the start as a smoothed-outcontinuum. The information we have about a continuum represents the microscopic information averaged over a volume. Classical thermodynamics is concerned only with continua.

1.2.2 The Concept of a ``System''


A thermodynamic system is a quantity of matter of fixed identity, around which we can draw a boundary (see Figure 1.3 for an example). The boundaries may be fixed or moveable. Work or heat can be transferred across the system boundary. Everything outside the boundary is the surroundings. When working with devices such as engines it is often useful to define the system to be an

identifiable volume with flow in and out. This is termed a control volume. An example is shown in Figure 1.5. A closed system is a special class of system with boundaries that matter cannot cross. Hence the principle of the conservation of mass is automatically satisfied whenever we employ a closed system analysis. This type of system is sometimes termed a control mass.

Figure 1.3: Piston (boundary) and gas (system)

Figure 1.4: Boundary around electric motor (system)

Figure 1.5: Sample control volume

1.2.3 The Concept of a ``State''


The thermodynamic state of a system is defined by specifying values of a set of measurable properties sufficient to determine all other properties. For fluid systems, typical properties are pressure, volume and temperature. More complex systems may require the specification of more unusual properties. As an example, the state of an electric battery requires the specification of the amount of electric charge it contains. Properties may be extensive or intensive. Extensive properties are additive. Thus, if the

system is divided into a number of sub-systems, the value of the property for the whole system is equal to the sum of the values for the parts. Volume is an extensive property. Intensive properties do not depend on the quantity of matter present. Temperature and pressure are intensive properties. Specific properties are extensive properties per unit mass and are denoted by lower case letters. For example:

Specific properties are intensive because they do not depend on the mass of the system.

The properties of a simple system are uniform throughout. In general, however, the properties of a system can vary from point to point. We can usually analyze a general system by sub-dividing it (either conceptually or in practice) into a number of simple systems in each of which the properties are assumed to be uniform. It is important to note that properties describe states only when the system is in equilibrium.

Muddy Points Specific properties (MP 1.1) What is the difference between extensive and intensive properties? (MP 1.2)

1.2.4 The Concept of ``Equilibrium''


The state of a system in which properties have definite, unchanged values as long as external conditions are unchanged is called an equilibrium state.

[Mechanical Equilibrium]

[Thermal

Equilibrium]

Figure 1.6: Equilibrium

A system in thermodynamic equilibrium satisfies:


1. mechanical equilibrium (no unbalanced forces) 2. thermal equilibrium (no temperature differences) 3. chemical equilibrium.

1.2.5 The Concept of a ``Process''


If the state of a system changes, then it is undergoing a process. The succession of states through which the system passes defines the path of the process. If, at the end of the process, the properties have returned to their original values, the system has undergone a cyclic process or a cycle. Note that even if a system has returned to its original state and completed a cycle, the state of the surroundings may have changed.

1.2.6 Quasi-Equilibrium Processes


We are often interested in charting thermodynamic processes between states on thermodynamic coordinates. Recall from the end of Section 1.2.3, however, that properties define a state only when a system is in equilibrium. If a process involves finite, unbalanced forces, the system can pass through non-equilibrium states, which we cannot treat. An extremely useful idealization, however, is that only ``infinitesimal'' unbalanced forces exist, so that the process can be viewed as taking place in a series of ``quasi-equilibrium'' states. (The term quasi can be taken to mean ``as if;'' you will see it used in a number of contexts such as quasi-one-dimensional, quasi-steady, etc.) For this to be true the process must be slow in relation to the time needed for the system to come to equilibrium internally. For a gas at conditions of interest to us, a given molecule can undergo roughly molecular collisions per second, so that, if ten collisions are needed to come to equilibrium, the equilibration time is on the order of seconds. This is generally much shorter than the time scales associated with the bulk properties of the flow (say the time needed for a fluid particle to move some significant fraction of the length of the device of interest). Over a large range of parameters, therefore, it is a very good approximation to view the thermodynamic processes as consisting of such a succession of equilibrium states, which we can chart. [VW, S& B: 2.3-2.4] The figures below demonstrate the use of thermodynamics coordinates to plot isolines, lines along which a property is constant. They include constant temperature lines, or isotherms, on a - diagram, constant volume lines, or isochors on a and constant pressure lines, or isobars, on a - diagram for an ideal gas. diagram,

Real substances may have phase changes (water to water vapor, or water to ice, for example), which we can also plot on thermodynamic coordinates. We will see such phase changes plotted and used for liquid-vapor power generation cycles in Chapter 8. A preview is given in Figure 1.15 at the end of this chapter.

Figure: Figure:

diagram diagram

Figure: - diagram Figure 1.7: Thermodynamics coordinates and isolines for an ideal gas

1.2.7 Equations of state


It is an experimental fact that two properties are needed to define the state of any pure substance in equilibrium or undergoing a steady or quasi-steady process. [VW, S & B: 3.1, 3.3]. Thus for a simple compressible gas like air,

where

is the volume per unit mass,

. In words, if we know

and

we know

, etc.

Any of these is equivalent to an equation , which is known as an equation of state. The equation of state for an ideal gas, which is a very good approximation to real gases at conditions that are typically of interest for aerospace applications 1.2, is

where .

is the volume per mol of gas and

is the ``Universal Gas Constant,''

A form of this equation which is more useful in fluid flow problems is obtained if we divide by the molecular weight, :

where R is

, which has a different value for different gases due to the different molecular .

weights. For air at room conditions,

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's All Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The First Law Contents Index

3.1 Some Properties of Engineering Cycles; Work and Efficiency


As preparation for our discussion of cycles (and as a foreshadowing of the second law), we examine two types of processes that concern interactions between heat and work. The first of these represents the conversion of work into heat. The second, which is much more useful, concerns the conversion of heat into work. The question we will pose is how efficient can this conversion be in the two cases.

Figure 3.1: Examples of the conversion of work into heat

Three examples of the first process are given in Figure 3.1. The first is the pulling of a block on a rough horizontal surface by a force which moves through some distance. Friction resists the pulling. After the force has moved through the distance, it is removed. The block then has no kinetic energy and the same potential energy it had when the force started to act. If we measured the temperature of the block and the surface we would find that it was higher than when we started. (High temperatures can be reached if the velocities of pulling are high; this is the basis of inertia welding.) The work done to move the block has been converted totally to heat. The second example concerns the stirring of a viscous liquid. There is work associated with the torque exerted on the shaft turning through an angle. When the stirring stops, the fluid comes to rest and there is (again) no change in kinetic or potential energy from the initial state. The fluid and the paddle wheels will be found to be hotter than when we started, however. The final example is the passage of a current through a resistance. This is a case of electrical work being converted to heat, indeed it models operation of an electrical heater. All the examples in Figure 3.1 have 100% conversion of work into heat. This 100% conversion could go on without limit as long as work were supplied. Is this true for the conversion of heat into work? To answer the last question, we need to have some basis for judging whether work is done in a given process. One way to do this is to ask whether we can construct a way that the process could result in the raising of a weight in a gravitational field. If so, we can say ``Work has been done.'' It may sometimes be difficult to make the link between a complicated thermodynamic process and the simple raising of a weight, but this is a rigorous test for the existence of work. One example of a process in which heat is converted to work is the isothermal (constant temperature) expansion of an ideal gas, as sketched in Figure 3.2. The system is the gas inside the chamber. As the gas expands, the piston does work on some external device. For an ideal gas, the internal energy is a function of temperature only, so that if the temperature is constant for some process the internal energy change is zero. To keep the temperature constant during the expansion, heat must be supplied. Because first law takes the form work. the

. This is a process that has 100% conversion of heat into

Figure 3.2: Isothermal expansion

The work exerted by the system is given by

where 1 and 2 denote the two states at the beginning and end of the process. The equation of state for an ideal gas is

with the number of moles of the gas contained in the chamber. Using the equation of state, the expression for work can be written as

(3..1)

For an isothermal process, , so that written in terms of the pressures at the beginning and end as

. The work can be

(3..2)

The lowest pressure to which we can expand and still receive work from the system is atmospheric pressure. Below this, we would have to do work on the system to pull the piston out further. There is thus a bound on the amount of work that can be obtained in the isothermal expansion; we cannot continue indefinitely. For a power or propulsion system, however, we would like a source of continuous power, in other words a device that would give power or propulsion as long as fuel was added to it. To do this, we need a series of processes where the system does not progress through a one-way transition from an initial state to a different final state, but rather cycles back to the initial state. What is looked for is in fact a thermodynamic cycle for the system. We define several quantities for a cycle:

is the heat absorbed by the system. is the heat rejected by the system.

is the net work done by the system.

The cycle returns to its initial state, so the overall energy change, , is zero. The net work done by the system is related to the magnitudes of the heat absorbed and the heat rejected by

The thermal efficiency of the cycle is the ratio of the work done to the heat absorbed. (Efficiencies are often usefully portrayed as ``What you get'' versus ``What you pay for.'' Here what we get is work and what we pay for is heat, or rather the fuel that generates the heat.) In terms of the heat absorbed and rejected, the thermal efficiency is

(3..3)

The thermal efficiency can only be 100% (complete conversion of heat into work) if ; a basic question is what is the maximum thermal efficiency for any arbitrary cycle? We examine this for several cases, including the Carnot cycle and the Brayton (or Joule) cycle, which is a model for the power cycle in a jet engine.

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal combustion Contents Index

3.6 Diesel Cycle

The Diesel cycle is a compression ignition (rather than spark ignition) engine. Fuel is sprayed into the cylinder at (high pressure) when the compression is complete, and there is ignition without a spark. An idealized Diesel engine cycle is shown in Figure 3.12.

Figure 3.12: The ideal Diesel cycle The thermal efficiency is given by:

This cycle can operate with a higher compression ratio than the Otto cycle because only air is compressed and there is no risk of auto-ignition of the fuel. Although for a given compression ratio the Otto cycle has higher efficiency, because the Diesel engine can be operated to higher compression ratio, the engine can actually have higher efficiency than an Otto cycle when both are operated at compression ratios that might be achieved in practice.

Muddy Points

When and where do we use ever ? (MP 3.8)

and

? Some definitions use

. Is it

Explanation of the above comparison between Diesel and Otto. (MP 3.9)

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal combustion Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.6 Diesel Cycle Up: 3. The First Law Previous: 3.4 Refrigerators and Heat Contents Index Subsections

3.5.1 Efficiency of an ideal Otto cycle 3.5.2 Engine work, rate of work per unit enthalpy flux

3.5 The Internal combustion engine (Otto Cycle)


[VW, S & B: 9.13]

The Otto cycle is a set of processes used by spark ignition internal combustion engines (2stroke or 4-stroke cycles). These engines a) ingest a mixture of fuel and air, b) compress it, c) cause it to react, thus effectively adding heat through converting chemical energy into thermal energy, d) expand the combustion products, and then e) eject the combustion products and replace them with a new charge of fuel and air. The different processes are shown in Figure 3.8:
1. Intake stroke, gasoline vapor and air drawn into engine ( 2. Compression stroke, , increase ( ). 3. Combustion (spark), short time, essentially constant volume ( absorbed from a series of reservoirs at temperatures 4. Power stroke: expansion ( ). 5. Valve exhaust: valve opens, gas escapes. to . ).

). Model: heat

6. ( ) Model: rejection of heat to series of reservoirs at temperatures to 7. Exhaust stroke, piston pushes remaining combustion products out of chamber (

. ).

We model the processes as all acting on a fixed mass of air contained in a piston-cylinder arrangement, as shown in Figure 3.10.

Figure 3.8: The ideal Otto cycle

Figure 3.9: Sketch of an actual Otto cycle

Figure 3.10: Piston and valves in a four-stroke internal combustion engine

The actual cycle does not have the sharp transitions between the different processes that the ideal cycle has, and might be as sketched in Figure 3.9.

3.5.1 Efficiency of an ideal Otto cycle


The starting point is the general expression for the thermal efficiency of a cycle:

The convention, as previously, is that heat exchange is positive if heat is flowing into the system or engine, so is negative. The heat absorbed occurs during combustion when the spark occurs, roughly at constant volume. The heat absorbed can be related to the temperature change from state 2 to state 3 as:

The heat rejected is given by (for a perfect gas with constant specific heats)

Substituting the expressions for the heat absorbed and rejected in the expression for thermal efficiency yields

We can simplify the above expression using the fact that the processes from 1 to 2 and from 3 to 4 are isentropic:

The quantity is called the compression ratio. In terms of compression ratio, the efficiency of an ideal Otto cycle is:

Figure 3.11: Ideal Otto cycle thermal efficiency

The ideal Otto cycle efficiency is shown as a function of the compression ratio in Figure 3.11. As the compression ratio, If , increases, increases, but so does .

is too high, the mixture will ignite without a spark (at the wrong location in the cycle).

3.5.2 Engine work, rate of work per unit enthalpy flux


The non-dimensional ratio of work done (the power) to the enthalpy flux through the engine is given by

There is often a desire to increase this quantity, because it means a smaller engine for the same power. The heat input is given by

where

is the heat of reaction, i.e. the chemical energy liberated per unit mass of fuel, is the fuel mass flow rate.

The non-dimensional power is

The quantities in this equation, evaluated at stoichiometric conditions are:

so

Muddy Points

How is

calculated? (MP 3.6)

What are ``stoichiometric conditions?'' (MP 3.7)

Next: 3.6 Diesel Cycle Up: 3. The First Law Previous: 3.4 Refrigerators and Heat Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 8.6 Enhancements of Rankine Up: 8. Power Cycles with Previous: 8.4 The Clausius-

Clapeyron Equation Contents Index

8.5 Rankine Power Cycles

Figure 8.11: Rankine power cycle with two-phase working fluid [Moran and Shapiro, Fundamentals of Engineering Thermodynamics] A schematic of the components of a Rankine cycle is shown in Figure 8.11. The cycle is shown on - , - , and - coordinates in Figure 8.12. The processes in the Rankine cycle are as follows:

1.

: Cold liquid at initial temperature is pressurized reversibly to a high pressure by a pump. In this process, the volume changes slightly. : Reversible constant pressure heating in a boiler to temperature .

2. 3.

: Heat added at constant temperature (constant pressure), with transition of liquid to vapor. 4. : Isentropic expansion through a turbine. The quality decreases from unity at point to 5. . : Liquid-vapor mixture condensed at temperature by extracting heat.

coordinates]

coordinates]

coordinates]

Figure 8.12: Rankine cycle diagram. Stations correspond to those in Figure 8.11

In the Rankine cycle, the mean temperature at which heat is supplied is less than the maximum temperature, , so that the efficiency is less than that of a Carnot cycle working between the same maximum and minimum temperatures. The heat absorption takes place at constant pressure over , but only the part is isothermal. The heat rejected occurs over ; this is at both constant temperature and pressure. To examine the efficiency of the Rankine cycle, we define a mean effective temperature, , in terms of the heat exchanged and the entropy differences:

The thermal efficiency of the cycle is

The compression and expansion processes are isentropic, so the entropy differences are related by

The thermal efficiency can be written in terms of the mean effective temperatures as

For the Rankine cycle, , . From this equation we see not only the reason that the cycle efficiency is less than that of a Carnot cycle, but the direction to move in terms of cycle design (increased ) if we wish to increase the efficiency.

There are several features that should be noted about Figure 8.12 and the Rankine cycle in

general:
1. The - and the - diagrams are not similar in shape, as they were with the perfect gas with constant specific heats. The slope of a constant pressure reversible heat addition line is, as derived in Chapter 6,

In the two-phase region, constant pressure means also constant temperature, so the slope of the constant pressure heat addition line is constant and the line is straight. 2. The effect of irreversibilities is represented by the dashed line from behavior during the expansion results in a value of entropy to . Irreversible

at the end state of the

expansion that is higher than . The enthalpy at the end of the expansion (the turbine exit) is thus higher for the irreversible process than for the reversible process, and, as seen for the Brayton cycle, the turbine work is thus lower in the irreversible case. 3. The Rankine cycle is less efficient than the Carnot cycle for given maximum and minimum temperatures, but, as said earlier, it is more effective as a practical power production device.

Muddy Points Where does degrees Rankine come from? Related to Rankine cycles? (MP 8.9)

Next: 8.6 Enhancements of Rankine Up: 8. Power Cycles with Previous: 8.4 The ClausiusClapeyron Equation Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal combustion Contents Index

3.6 Diesel Cycle


The Diesel cycle is a compression ignition (rather than spark ignition) engine. Fuel is sprayed into the cylinder at (high pressure) when the compression is complete, and there is ignition without a spark. An idealized Diesel engine cycle is shown in Figure 3.12.

Figure 3.12: The ideal Diesel cycle The thermal efficiency is given by:

This cycle can operate with a higher compression ratio than the Otto cycle because only air is compressed and there is no risk of auto-ignition of the fuel. Although for a given compression ratio the Otto cycle has higher efficiency, because the Diesel engine can be operated to higher compression ratio, the engine can actually have higher efficiency than an Otto cycle when both are operated at compression ratios that might be achieved in practice.

Muddy Points

When and where do we use ever ? (MP 3.8)

and

? Some definitions use

. Is it

Explanation of the above comparison between Diesel and Otto. (MP 3.9)

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal combustion Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index Subsections

11.6.1 Notation and station numbering 11.6.2 Ideal Assumptions 11.6.3 Ideal Ramjet o 11.6.3.1 Thrust o 11.6.3.2 Fuel Air Ratio
o o o

11.6.3.3 Specific impulse, 11.6.3.4 Representative performance values 11.6.3.5 Recapitulation

11.6.4 Turbojet Engine 11.6.5 Effect of Departures from Ideal Behavior o 11.6.5.1 Parameters reflecting design choices o 11.6.5.2 Parameters reflecting the ability to design and execute efficient components

11.6 Performance of Jet Engines


In Chapter 3 we represented a gas turbine engine using a Brayton cycle and derived expressions for efficiency and work as functions of the temperature at various points in the cycle. In this section we will perform further ideal cycle analysis to express the thrust and fuel efficiency of engines in terms of useful design variables, including design limits, flight

conditions, and design choices. The expressions we develop will allow us to define a particular mission and then determine the optimum component characteristics (e.g. compressor, combustor, turbine) for an engine for a given mission. Note that ideal cycle analysis addresses only the thermodynamics of the airflow within the engine. It does not describe the details of the components (the blading, the rotational speed, etc.), but only the results the various components produce (e.g. pressure ratios, temperature ratios). In Chapter 12 we will look in greater detail at how some of the components (the turbine and the compressor) produce these effects.

11.6.1 Notation and station numbering


Notation:

where the capital subscript

will refer to the turbine. Stagnation properties,

and

, are more

easily measured quantities than static properties ( and ). Thus, it is standard convention to express the performance of various components in terms of stagnation pressure and temperature ratios:

total or stagnation pressure ratio across component ( , total or stagnation temperature ratio across component (

, , ,

, ,

, ,

, ,

) ),

where , , , , and engine we will use the numbers in Figure 11.12.

, . To identify individual stations within the

Figure 11.12: Gas turbine engine station numbering.

11.6.2 Ideal Assumptions


1. Inlet/Diffuser: , (adiabatic, isentropic)

2. Compressor or fan: 3. Combustor/burner or afterburner:

, ,

4. Turbine: 5. Nozzle: , .

11.6.3 Ideal Ramjet


To get started with a simple example (no turbomachinery), we will reexamine the ideal ramjet, picking up where we left off in Section 3.7.3. (Note that we will continue to use station 5 as the exit station, consistent with Figure 3.25.)

11.6.3.1 Thrust
The coordinate system and control volume are chosen to be fixed to the ramjet. The thrust, , is given by:

where and are the exit and inlet flow velocities, respectively. The thrust can be put in terms of nondimensional parameters as follows:

Using

in the expression for stagnation pressure,

The ratios of stagnation pressure to static pressure at inlet and exit of the ramjet are

The ratios of stagnation to static pressure at exit and at inlet are the same, with the consequence that the inlet and exit Mach numbers are also the same.

To find the thrust we need to find the ratio of the temperature at exit and the temperature at inlet. This is given by

where

is the stagnation temperature ratio across the combustor (burner). The thrust is thus

11.6.3.2 Fuel Air Ratio


To find the Isp will will need to find the ramjet fuel-air ratio, burner, as shown in Figure 11.13, we get: . Using a control volume around the

From the steady flow energy equation:

The exit mass flow is not greatly different from the inlet mass flow, , because the fuel-air ratio is much less than unity (generally several percent). We thus neglect the difference between the mass flows and obtain

with

Thus, the fuel-air ratio is

The fuel-air ratio,

, depends on the fuel properties (

), the desired flight parameters ( ).

), the ramjet performance (

), and the temperature of the atmosphere (

Figure 11.13: Control volume over the burner

11.6.3.3 Specific impulse,


The specific impulse for the ramjet is given by

The specific impulse can be written in terms of fuel properties and flight and vehicle characteristics as

We wish to explore the parameter dependency of the above expression, which is a complicated formula. How can we do this? What are the important effects of the different parameters? How do we best capture the ramjet performance behavior? To make effective comparisons, we need to add some additional information concerning the operational behavior. An important case to examine is when we have stoichiometric conditions and all the fuel burns (denoted by ). What happens in this situation as the

flight Mach number, , increases? is fixed so increases, but the maximum temperature does not increase much because of dissociation: the reaction does not go to completion at high temperature. A useful approximation is therefore to take constant

for stoichiometric operation. In the stratosphere, from 10 to 30 km, . The maximum temperature ratio is

For the stoichiometric ramjet,

Using the expression for

, the specific impulse is

11.6.3.4 Representative performance values

Figure 11.14: Thrust per unit mass flow and specific impulse for ideal ramjet with stoichiometric combustion [Kerrebrock]

A plot of the performance of the stoichiometric ramjet is shown in Figure 11.14. The figure shows that for the parameters used, the best operating range of a hydrocarbon-fueled ramjet is stratosphere, . The parameters used are for hydrocarbons, such that , in the .

11.6.3.5 Recapitulation

Between Section 3.7.3 and this section we have:


Examined the Brayton Cycle for ramjet propulsion, Found Found Examined as a function of , and . , and

and the relation between and as a function of

Muddy Points What exactly is the specific impulse, How is found for rockets in space where rather than

, a measure of? (MP 11.1) ? (MP 11.2)

Why does industry use

? Is there an advantage to this? (MP 11.3)

Why isn't mechanical efficiency an issue with ramjets? (MP 11.4) How is thrust created in a ramjet? (MP 11.5)

What is the relation between of ? (MP 11.6)

and the existence of the maximum value

Why didn't we have a 2s point for the Brayton cycle with non-ideal components? (MP 11.7)

What is the variable

? (MP 11.8)

11.6.4 Turbojet Engine

Figure 11.15: Schematic with appropriate component notations added.

We now examine an engine with turbomachinery, as shown in Figure 11.15. Methodology:

1. Find thrust by finding in terms of , temperature ratios, etc. 2. Use a power balance to relate turbine parameters to compressor parameters 3. Use an energy balance across the combustor to relate the combustor temperature rise to the fuel flow rate and fuel energy content. First write-out the expressions for thrust and Isp:

where

is the fuel/air mass flow ratio,

and

Now we have to do a little algebra to manipulate these expressions into more useful forms. First we write an expression for the exit velocity:

Noting that

we can write

Thus (11..1)

which expresses the exit temperature as a function of the inlet temperature, the Mach number, and the temperature changes across each component. We now write the pressure at the exit in a similar manner:

Since

and

we write

and then equate this to our expression for the temperature in (11.1):

and

We now continue on the path to our expression for

Therefore

Now we have two steps left. First we write in terms of , by noting that they are related by the condition that the power used by the compressor is equal to the power extracted by the turbine. Second, we put the burner temperature ratio in terms of the exit temperature of the burner, ( more specifically ) since this is the hottest point in the engine and is a frequent benchmark used for judging various designs. or

The steady flow energy equation states

Assuming that the compressor and turbine are adiabatic, then

Since the turbine shaft is connected to the compressor shaft

Assuming the mass flow and specific heats are the same between compressor and turbine, this can be rewritten as

where

so

or

That was the first step relating the temperature rise across the turbine to that across the compressor. The remaining step is to write the temperature rise across the combustor in terms of .

and for an engine with an afterburner

Now substituting our expressions for , and the first expression we wrote for thrust, we get:

into our expression for

, and finally into

This is what we were seeking, an expression for thrust in terms of important design parameters and flight parameters:

By adding and substracting the quantity

we may write this write in another form which is often used,

A recap of the important variables:

Our final step involves writing the specific impulse and other measures of efficiency in terms of these same parameters. We begin by writing the First Law across the combustor to relate the fuel flow rate and heating value of the fuel to the total enthalpy rise.

and

where again,

is the fuel/air mass flow ratio. The specific impulse becomes

where is expressed in terms of typical design parameters, flight conditions, and physical constants

Similarly, we can write our overall efficiency as

or

Using the definition from before, the ideal thermal efficiency is

and the propulsive efficiency can be found as

We can now use these equations to better understand the performance of a simple turbojet engine. We will use the following parameters (with ):

Table 11.1: Turbojet Mission Parameters Mach number Altitude Ambient Temp. Speed of sound 0 0.85 2.0 Sea level 12 km 18 km 288 K 217 K 217 K 340 m/s 295 m/s 295 m/s

Note it is more typical to work with the compressor pressure ratio ( ratio ( ) so we will substitute the isentropic relationship:

) rather than the temperature

into the equations before plotting the results in Figures 11.16, 11.17, and 11.18.

Figure 11.16: Performance of an ideal turbojet engine as a function of flight Mach number.

Figure 11.17: Performance of an ideal turbojet engine as a function of compressor pressure ratio and flight Mach number.

Figure 11.18: Performance of an ideal turbojet engine as a function of compressor pressure ratio and turbine inlet temperature.

11.6.5 Effect of Departures from Ideal Behavior -- Real Cycle behavior


To conclude this chapter, we will now improve our estimates of cycle performance by including the effects of irreversibility. We will use the Brayton cycle as an example. What are the sources of non-ideal performance and departures from reversibility?

Losses (entropy production) in the compressor and the turbine. Stagnation pressure decrease in the combustor. Heat transfer.

We take into account here only irreversibility in the compressor and in the turbine. Because

of these irreversibilities, we need more work, (the changes in kinetic energy from inlet to exit of the compressor are neglected), to drive the compressor than in the ideal situation. We also get less work, , back from the turbine. The consequence, as can be inferred from Figure 11.19, is that the net work from the engine is less than in the cycle with ideal components.

Figure 11.19: Gas turbine engine (Brayton) cycle showing effect of departure from ideal behavior in compressor and turbine

To develop a quantitative description of the effect of these departures from reversible behavior, consider a perfect gas with constant specific heats and neglect kinetic energy at the inlet and exit of the turbine and compressor. We define the turbine adiabatic efficiency as

where is specified to be at the same pressure ratio as the compressor, the compressor adiabatic efficiency:

. There is a similar metric for

again for the same pressure ratio. Note that for the turbine the ratio is the actual work delivered divided by the ideal work, whereas for the compressor the ratio is the ideal work needed divided by the actual work required. These are not thermal efficiencies, but rather measures of the degree to which the compression and expansion approach the ideal processes.

We now wish to find the net work done in the cycle and the efficiency. The net work is given

either by the difference between the heat received and rejected or the work of the compressor and turbine, where the convention is that heat received is positive and heat rejected is negative and work done is positive and work absorbed is negative.

The thermal efficiency is

We need to calculate

From the definition of

With

Similarly, by the definition

we can find

The thermal efficiency can now be found:

With

and

or

There are several non-dimensional parameters that appear in this expression for thermal efficiency. We list these in the two sections below and show their effects in accompanying figures.

11.6.5.1 Parameters reflecting design choices

11.6.5.2 Parameters reflecting the ability to design and execute efficient components

In addition to efficiency, net rate of work is a quantity we need to examine,

Putting this in a non-dimensional form,

[Non-dimensional work as a function of cycle pressure ratio for different values of turbine entry

temperature divided by compressor entry temperature] [Overall cycle efficiency as a function of pressure ratio for different values of turbine entry temperature

divided by compressor entry temperature] [Overall cycle efficiency as a function of cycle pressure ratio for different component

efficiencies] Figure 11.20: Non-dimensional power and efficiency for a non-ideal gas turbine engine [from Cumpsty, Jet Propulsion]

Trends in net power and efficiency are shown in Figure 11.20 for parameters typical of advanced civil engines. Some points to note in the figure:

For any , the optimum pressure ratio for maximum is not the highest that can be achieved, as it is for the ideal Brayton cycle. The ideal analysis is too idealized in this regard. The highest efficiency also occurs closer to the pressure ratio for maximum power than in the case of an ideal cycle. Choosing this as a design criterion will therefore not lead to the efficiency penalty inferred from ideal cycle analysis. There is a strong sensitivity to the component efficiencies. For example, for , the cycle efficiency is roughly two-thirds of the ideal value. or pressure ratio less than that for

The maximum power occurs at a value of max

(this trend is captured by ideal analysis). are strongly dependent on the maximum

The maximum power and maximum temperature, .

A scale diagram of a Brayton cycle with non-ideal compressor and turbine behaviors, in terms of temperature-entropy ( - ) and pressure-volume ( - ) coordinates is given below as Figure 11.21.

[]

[]

Figure 11.21: Scale diagram of non-ideal gas turbine cycle. Nomenclature is shown in the figure. Pressure ratio , , , compressor and turbine efficiencies Cumpsty, Jet Propulsion] [from

Muddy Points Isn't it possible for the mixing of two gases to go from the final state to the initial state? If you have two gases in a box, they should eventually separate by density, right? (MP 11.9)

How can

be the maximum turbine inlet temperature? (MP 11.10)

When there are losses in the turbine that shift the expansion in - diagram to the right, does this mean there is more work than ideal since the area is greater? (MP 11.11) For an afterburning engine, why must the nozzle throat area increase if the temperature of the fluid is increased? (MP 11.12) Why doesn't the pressure in the afterburner go up if heat is added? (MP 11.13) Why is the flow in the nozzle choked? (MP 11.14) What's the point of having a throat if it creates a retarding force? (MP 11.15) Why isn't the stagnation temperature conserved in this steady flow? (MP 11.16)

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 17.2 Combined Conduction and Up: 17. Convective Heat Transfer Previous: 17. Convective Heat Transfer Contents Index

17.1 The Reynolds Analogy


We describe the physical mechanism for the heat transfer coefficient in a turbulent boundary layer because most aerospace vehicle applications have turbulent boundary layers. The treatment closely follows that in Eckert and Drake (1959). Very near the wall, the fluid motion is smooth and laminar, and molecular conduction and shear are important. The shear stress, , at a plane is given by (where is the dynamic viscosity), and the heat flux by . The latter is the same expression that was used for a solid. The boundary layer is a region in which the velocity is lower than the free stream as shown in Figures 17.2 and 17.3. In a turbulent boundary layer, the dominant mechanisms of shear stress and heat transfer change in nature as one moves away from the wall.

Figure 17.3: Velocity profile near a surface

As one moves away from the wall (but still in the boundary layer), the flow is turbulent. The fluid particles move in random directions and the transfer of momentum and energy is mainly through interchange of fluid particles, shown schematically in Figure 17.4.

Figure 17.4: Momentum and energy exchanges in turbulent flow

With reference to Figure 17.4, because of the turbulent velocity field, a fluid mass penetrates the plane per unit time and unit area. In steady flow, the same amount crosses from the other side. Fluid moving up transports heat downwards. If . Fluid moving

down transports flow , given by

, there is a turbulent downwards heat , that results. and fluid moving down has momentum .

Fluid moving up also has momentum

The net flux of momentum down per unit area and time is therefore . This net flux of momentum per unit area and time is a force per unit area or stress, given by
(17..3)

Based on these considerations, the relation between heat flux and shear stress at plane

is (17..4)

or (again approximately) (17..5)

since the locations of planes 1-1 and 2-2 are arbitrary.

For the laminar region, the heat flux towards the wall is expression for the shear stress, , yields

and dividing by the

(17..6)

The same relationship is applicable in laminar or turbulent flow if differently,

or, expressed slightly

(17..7)

where

is the kinematic viscosity, and

is the thermal diffusivity.

The quantity is known as the Prandtl number ( ), after the man who first presented the idea of the boundary layer and was one of the pioneers of modern fluid mechanics. For gases, Prandtl numbers are in fact close to unity and for air at room temperature. The Prandtl number varies little over a wide range of temperatures: approximately 3% from 300-2000 K. We want a relation between the values at the wall (at which and ) and those in the free stream. To get this, we integrate the expression for from the wall to the free stream

(17..8)

where the relation between heat transfer and shear stress has been taken as the same for both the laminar and the turbulent portions of the boundary layer. The assumption being made is that the mechanisms of heat and momentum transfer are similar. Equation (17.8) can be integrated from the wall to the freestream (conditions ``at ''):

(17..9)

where

and

are assumed constant.

Carrying out the integration yields

(17..10)

where

is the velocity and

is the specific heat. In Equation (17.10),

is the heat flux to the

wall and is the shear stress at the wall. The relation between skin friction (shear stress) at the wall and heat transfer is thus (17..11)

The quantity

is known as the skin friction coefficient and is denoted by . The skin friction coefficient has been tabulated (or computed) for a large number of situations. If we define a non-dimensional quantity (17..12)

known as the Stanton Number, we can write an expression for the heat transfer coefficient,

as

(17..13)

Equation (17.13) provides a useful estimate of , or , based on knowing the skin friction, or drag. The direct relationship between the Stanton Number and the skin friction coefficient is

(17..14)

The relation between the heat transfer and the skin friction coefficient (17..15)

is known as the Reynolds analogy between shear stress and heat transfer. The Reynolds analogy is extremely useful in obtaining a first approximation for heat transfer in situations in which the shear stress is ``known.''

An example of the use of the Reynolds analogy is in analysis of a heat exchanger. One type of heat exchanger has an array of tubes with one fluid flowing inside and another fluid flowing outside, with the objective of transferring heat between them. To begin, we need to examine the flow resistance of a tube. For fully developed flow in a tube, it is more appropriate to use an average velocity and a bulk temperature approximate relation for the heat transfer is . Thus, an

(17..16)

The fluid resistance (drag) is all due to shear forces and is given by the tube ``wetted'' area (perimeter length). The total heat transfer, , is

, where , so that

is

(17..17)

The power, velocity,

, to drive the flow through a resistance is given by the product of the drag and the , so that

(17..18)

The mass flow rate is given by

, where

is the cross sectional area. For a given mass or as , i.e.,

flow rate and overall heat transfer rate, the power scales as

(17..19)

Equations (17.18) and (17.19) show that to decrease the power dissipated, we need to decrease , which can be accomplished by increasing the cross-sectional area. Two possible heat exchanger configurations are sketched in Figure 17.5; the one on the right will have a lower loss.

Figure 17.5: Heat exchanger configurations

To recap, there is an approximate relation between skin friction (momentum flux to the wall) and heat transfer called the Reynolds analogy that provides a useful way to estimate heat transfer rates in situations in which the skin friction is known. The relation is expressed by

or

or

The Reynolds analogy can be used to give information about scaling of various effects as well as initial estimates for heat transfer. It is emphasized that it is a useful tool based on a hypothesis about the mechanism of heat transfer and shear stress and not a physical law.

Muddy Points What is the ``analogy'' that we are discussing? Is it that the equations are similar? (MP 17.2) In what situations does the Reynolds analogy ``not work?'' (MP 17.3)

Next: 17.2 Combined Conduction and Up: 17. Convective Heat Transfer Previous: 17. Convective Heat Transfer Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.8 Some Overall Comments Up: 6. Applications of the Previous: 6.6 Entropy and Unavailable Contents Index

6.7 Examples of Lost Work in Engineering Processes


1. Lost work in Adiabatic Throttling: Entropy and Stagnation Pressure Changes

Figure 6.8: Adiabatic Throttling

2. A process we have encountered before is adiabatic throttling of a gas, by a valve or other device as shown in Figure 6.8. The velocity is denoted by . There is no shaft work and no heat transfer and the flow is steady. Under these conditions we can use the first law for a control volume (the Steady Flow Energy Equation) to make a statement about the conditions upstream and downstream of the valve:

3. 4. where is the stagnation enthalpy, corresponding to a (possibly fictitious) state with zero , even if

velocity. The stagnation enthalpy is the same at stations 1 and 2 if the flow processes are not reversible.

5. For a perfect gas with constant specific heats, and relation between the static and stagnation temperatures is:

. The

6.

7. where is the speed of sound and is the Mach number, . In deriving this result, use has only been made of the first law, the equation of state, the speed of sound, and the definition of the Mach number. Nothing has yet been specified about whether the process of stagnating the fluid is reversible or irreversible.

8. When we define the stagnation pressure, however, we do it with respect to isentropic deceleration to the zero velocity state. For an isentropic process

9.

10. The relation between static and stagnation pressures is

11.

Figure 6.9: Static and stagnation pressures and temperatures

12. The stagnation state is defined by

. In

addition, . The static and stagnation states are shown in - coordinates in Figure 6.9. 13. Stagnation pressure is a key variable in propulsion and power systems. To see why, we examine the relation between stagnation pressure, stagnation temperature, and entropy. The form of the combined first and second law that uses enthalpy is

(6..8)

14.

Figure 6.10:Stagnation and static states

15. This holds for small changes between any thermodynamic states and we can apply it to a situation in which we consider differences between stagnation states, say one

state having properties

and the other having

properties (see Figure 6.10). The corresponding static states are also indicated. Because the entropy is the same at static and stagnation conditions, needs no subscript. Writing (6.8) in terms of stagnation conditions yields

16. 17. Both sides of the above are perfect differentials and can be integrated as

18.

19. For a process with , the stagnation enthalpy, and hence the stagnation temperature, is constant. In this situation, the stagnation pressure is related directly to the entropy as,

(6..9)

20.

Figure 6.11:Losses reflected in changes in stagnation pressure when

21. Figure 6.11 shows this relation on a - diagram. We have seen that the entropy is related to the loss, or irreversibility. The stagnation pressure plays the role of an indicator of loss if the stagnation temperature is constant. The utility is that it is the stagnation pressure (and temperature) which are directly measured, not the entropy. The throttling process is a representation of flow through inlets, nozzles, stationary turbomachinery blades, and the use of stagnation pressure as a measure of loss is a practice that has widespread application. 22. Equation (6.9) can be put in several useful approximate forms. First, we note that for aerospace applications we are (hopefully!) concerned with low loss devices, so that the stagnation pressure change is small compared to the inlet level of stagnation pressure,

23. 24. Expanding the logarithm (using ),

26. or

25.

27. 28. Another useful form is obtained by dividing both sides by and taking the limiting forms ). Doing

of the expression for stagnation pressure in the limit of low Mach number ( this, we find:

29. 30. The quantity on the right can be interpreted as the change in the ``Bernoulli constant'' for incompressible (low Mach number) flow. The quantity on the left is a non-dimensional entropy change parameter, with the term now representing the loss of mechanical energy associated with the change in stagnation pressure.

31. To summarize:
1. For many applications the stagnation temperature is constant and the change in stagnation pressure is a direct measure of the entropy increase. 2. Stagnation pressure is the quantity that is actually measured so that linking it to entropy (which is not measured) is useful. 3. We can regard the throttling process as a ``free expansion'' at constant temperature from the initial stagnation pressure to the final stagnation pressure. We thus know that, for the process, the work we need to do to bring the gas back to the initial state is , which is the ``lost work'' per unit mass.

Muddy Points Why do we find stagnation enthalpy if the velocity never equals zero in the flow? (MP 6.13) Why does remain constant for throttling? (MP 6.14)

32. Adiabatic Efficiency of a Propulsion System Component (Turbine)

Figure 6.12: Schematic of turbine and associated thermodynamic representation in coordinates

33. A schematic of a turbine and the accompanying thermodynamic diagram are given in Figure 6.12. There is a pressure and temperature drop through the turbine and it produces work. There is no heat transfer so the expressions that describe the overall shaft work and the shaft work per unit mass are
(6..10)

(6..11)

34. 35. If the difference in the kinetic energy at inlet and outlet can be neglected, Equation (6.11) reduces to 36. 37. The adiabatic efficiency of the turbine is defined as

38. 39. The performance of the turbine can be represented in an - plane (similar to a plane for a perfect gas with constant specific heats) as shown in Figure 6.12. From the figure the adiabatic efficiency is

40.

41. The adiabatic efficiency can therefore be written as

42. 43. The non-dimensional term represents the departure from isentropic (reversible) processes and hence a loss. The quantity is the enthalpy difference for two states along a constant pressure line (see diagram). From the combined first and second laws, for a constant pressure process, small changes in enthalpy are related to the entropy change by , or approximately, 44.

45. The adiabatic efficiency can thus be approximated as

47. The quantity irreversibility.

46. represents a useful figure of merit for fluid machinery inefficiency due to

48.

Muddy Points 49. How do you tell the difference between shaft work/power and flow work in a turbine, both conceptually and mathematically? (MP 6.15)
50. Isothermal Expansion with Friction

Figure 6.13: Isothermal expansion with friction

51. In a more general look at the isothermal expansion, we now drop the restriction to frictionless processes. As seen in Figure 6.13, work is done to overcome friction. If the kinetic energy of the piston is negligible, a balance of forces tells us that
52. 53. During the expansion, the piston and the walls of the container will heat up because of the friction. The heat will be (eventually) transferred to the atmosphere; all frictional work ends up as heat transferred to the surrounding atmosphere. 54.

55. The amount of heat transferred to the atmosphere due to the frictional work only is thus,

56.

57. The entropy change of the atmosphere (considered as a heat reservoir) due to the

frictional work is 58. The engine operates in a cycle and the entropy change for the complete cycle is zero

(because entropy is a state variable). Therefore,

59. 60. The total entropy change is

61.

62. Suppose we had an ideal reversible engine working between these same two temperatures, which extracted the same amount of heat, temperature reservoir, and rejected heat of magnitude reservoir. The work done by this reversible engine is
63. 64. For the reversible engine the total entropy change over a cycle is

, from the high to the low temperature

65.

66. Combining the expressions for work and for the entropy changes,
67. 68. The entropy change for the irreversible cycle can therefore be written as

69. 70. The difference in work that the two cycles produce is proportional to the entropy that is generated during the cycle: 71. 72. The second law states that the total entropy generated is greater than zero for an irreversible process, so that the reversible work is greater than the actual work of the irreversible cycle.

73. An ``engine effectiveness,'' , can be defined as the ratio of the actual work obtained divided by the work that would have been delivered by a reversible engine operating between the two temperatures and :

74. 75. The departure from a reversible process is directly reflected in the entropy change and the decrease in engine effectiveness.

76.

Muddy Points 77. Why does ? (MP 6.17) 78. In discussing the terms ``closed system'' and ``isolated system,'' can you assume that you are discussing a cycle or not? (MP 6.18) 79. Does a cycle process have to have ? (MP 6.19) 80. In a real heat engine, with friction and losses, why is ? (MP 6.20)
81. Propulsive Power and Entropy Flux

still 0 if

The final example in this section combines a number of ideas presented in this subject and in Unified in the development of a relation between entropy generation and power needed to propel a vehicle. Figure 6.15 shows an aerodynamic shape (airfoil) moving through the atmosphere at a constant velocity. A coordinate system fixed to the vehicle has been adopted so that we see the airfoil as fixed and the air far away from the airfoil moving at a velocity . Streamlines of the flow have been sketched, as has the velocity distribution at station ``0'' far upstream and station ``d'' far downstream. The airfoil has a wake, which mixes with the surrounding air and grows in the downstream direction. The extent of the wake is also indicated. Because of the lower velocity in the wake the area between the stream surfaces is larger downstream than upstream.

Figure 6.15: Airfoil with wake and control volume for analysis of propulsive power requirement

We use a control volume description and take the control surface to be defined by the two stream surfaces and two planes at station 0 and station d. This is useful in

simplifying the analysis because there is no flow across the stream surfaces. The area of the downstream plane control surface is broken into , which is the area

outside the wake and , which is the area occupied by wake fluid, i.e., fluid that has suffered viscous losses. The control surface is also taken far enough away from the vehicle so that the static pressure can be considered uniform. For fluid which is not in the wake (no viscous forces), the momentum equation is

Uniform static pressure therefore implies uniform velocity, so that on

the velocity is

equal to the upstream value, . The downstream velocity profile is actually continuous, as indicated. It is approximated in the analysis as a step change to make the algebra a bit simpler. (The conclusions apply to the more general velocity profile as well and we would just need to use integrals over the wake instead of the algebraic expressions below.)

The equation expressing mass conservation for the control volume is


(6..12)

The vertical face of the control surface is far downstream of the body. By this station, the wake fluid has had much time to mix and the velocity in the wake is close to the free stream value, . We can thus write, (6..13)

(We chose our control surface so the condition

was upheld.)

The integral momentum equation (control volume form of the momentum equation) can be used to find the drag on the vehicle:
(6..14)

There is no pressure contribution in Eq. (6.14) because the static pressure on the control surface is uniform. Using the form given for the wake velocity and expanding the terms in the momentum equation we obtain
(6..15)

The last term in the right hand side of the momentum equation, , is small by virtue of the choice of control surface and we can neglect it. Doing this and grouping the terms on the right hand side of Eq. (6.15) in a different manner, we have

The terms in the square brackets on both sides of this equation are the continuity equation multiplied by . They thus sum to zero leaving the curly bracketed terms as (6..16)

The wake mass flow is . All this flow has a velocity defect (compared to the free stream) of , so that the defect in flux of momentum (the mass flow in the wake times the velocity defect) is, to first order in ,

The combined first and second law gives us a means of relating the entropy and velocity:

The pressure is uniform ( ) at the downstream station. There is no net shaft work or heat transfer to the wake so that the mass flux of stagnation enthalpy is constant. We can also approximate that the condition of constant stagnation enthalpy holds locally on all

streamlines. Applying both of these to the combined first and second law yields

For the present situation,

, so that (6..17)

In Equation (6.17) the upstream temperature is used because differences between wake quantities and upstream quantities are small at the downstream control station. The entropy can be related to the drag as (6..18)

The quantity is the entropy flux (mass flux times the entropy increase per unit mass; in the general case we would express this by an integral over the locally varying wake velocity and density). The power needed to propel the vehicle is the product of , . From Eq. (6.18), this can be related to the entropy flux in the wake to yield a compact expression for the propulsive power needed in terms of the wake entropy flux: (6..19)

This amount of work is dissipated per unit time in connection with sustaining the vehicle motion. Equation (6.19) is another demonstration of the relation between lost work and entropy generation, in this case manifested as power that needs to be supplied because of dissipation in the wake.

Muddy Points

Is it safe to say that entropy is the tendency for a system to go into disorder? (MP 6.21)

Next: 6.8 Some Overall Comments Up: 6. Applications of the Previous: 6.6 Entropy and Unavailable Contents Index

Thermodynamics and Propulsion

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index Subsections

11.6.1 Notation and station numbering 11.6.2 Ideal Assumptions 11.6.3 Ideal Ramjet o 11.6.3.1 Thrust o 11.6.3.2 Fuel Air Ratio
o o o

11.6.3.3 Specific impulse, 11.6.3.4 Representative performance values 11.6.3.5 Recapitulation

11.6.4 Turbojet Engine 11.6.5 Effect of Departures from Ideal Behavior o 11.6.5.1 Parameters reflecting design choices o 11.6.5.2 Parameters reflecting the ability to design and execute efficient components

11.6 Performance of Jet Engines


In Chapter 3 we represented a gas turbine engine using a Brayton cycle and derived expressions for efficiency and work as functions of the temperature at various points in the cycle. In this section we will perform further ideal cycle analysis to express the thrust and fuel efficiency of engines in terms of useful design variables, including design limits, flight conditions, and design choices.

The expressions we develop will allow us to define a particular mission and then determine the optimum component characteristics (e.g. compressor, combustor, turbine) for an engine for a given mission. Note that ideal cycle analysis addresses only the thermodynamics of the airflow within the engine. It does not describe the details of the components (the blading, the rotational speed, etc.), but only the results the various components produce (e.g. pressure ratios, temperature ratios). In Chapter 12 we will look in greater detail at how some of the components (the turbine and the compressor) produce these effects.

11.6.1 Notation and station numbering


Notation:

where the capital subscript

will refer to the turbine. Stagnation properties, and

and

, are

more easily measured quantities than static properties (

). Thus, it is standard

convention to express the performance of various components in terms of stagnation pressure and temperature ratios:

total or stagnation pressure ratio across component ( , total or stagnation temperature ratio across component (

, , ,

, ,

, ,

, ,

) ),

where , , , , and engine we will use the numbers in Figure 11.12.

, . To identify individual stations within the

Figure 11.12: Gas turbine engine station numbering.

11.6.2 Ideal Assumptions


1. Inlet/Diffuser: , (adiabatic, isentropic)

2. Compressor or fan: 3. Combustor/burner or afterburner:

, ,

4. Turbine: 5. Nozzle: , .

11.6.3 Ideal Ramjet


To get started with a simple example (no turbomachinery), we will reexamine the ideal ramjet, picking up where we left off in Section 3.7.3. (Note that we will continue to use station 5 as the exit station, consistent with Figure 3.25.)

11.6.3.1 Thrust
The coordinate system and control volume are chosen to be fixed to the ramjet. The thrust, , is given by:

where and are the exit and inlet flow velocities, respectively. The thrust can be put in terms of nondimensional parameters as follows:

Using

in the expression for stagnation pressure,

The ratios of stagnation pressure to static pressure at inlet and exit of the ramjet are

The ratios of stagnation to static pressure at exit and at inlet are the same, with the consequence that the inlet and exit Mach numbers are also the same.

To find the thrust we need to find the ratio of the temperature at exit and the temperature at inlet. This is given by

where

is the stagnation temperature ratio across the combustor (burner). The thrust is thus

11.6.3.2 Fuel Air Ratio


To find the Isp will will need to find the ramjet fuel-air ratio, burner, as shown in Figure 11.13, we get: . Using a control volume around the

From the steady flow energy equation:

The exit mass flow is not greatly different from the inlet mass flow, , because the fuel-air ratio is much less than unity (generally several percent). We thus neglect the difference between the mass flows and obtain

with

Thus, the fuel-air ratio is

The fuel-air ratio,

, depends on the fuel properties (

), the desired flight parameters ( ).

), the ramjet performance (

), and the temperature of the atmosphere (

Figure 11.13: Control volume over the burner

11.6.3.3 Specific impulse,


The specific impulse for the ramjet is given by

The specific impulse can be written in terms of fuel properties and flight and vehicle characteristics as

We wish to explore the parameter dependency of the above expression, which is a complicated formula. How can we do this? What are the important effects of the different parameters? How do we best capture the ramjet performance behavior? To make effective comparisons, we need to add some additional information concerning the operational behavior. An important case to examine is when we have stoichiometric conditions and all the fuel burns (denoted by ). What happens in this situation as

the flight Mach number, , increases? is fixed so increases, but the maximum temperature does not increase much because of dissociation: the reaction does not go to completion at high temperature. A useful approximation is therefore to take for stoichiometric operation. In the stratosphere, from 10 to 30 constant

km,

. The maximum temperature ratio is

For the stoichiometric ramjet,

Using the expression for

, the specific impulse is

11.6.3.4 Representative performance values

Figure 11.14: Thrust per unit mass flow and specific impulse for ideal ramjet with stoichiometric combustion [Kerrebrock]

A plot of the performance of the stoichiometric ramjet is shown in Figure 11.14. The figure shows that for the parameters used, the best operating range of a hydrocarbon-fueled ramjet is stratosphere, . The parameters used are for hydrocarbons, such that , in the .

11.6.3.5 Recapitulation
Between Section 3.7.3 and this section we have:

Examined the Brayton Cycle for ramjet propulsion, Found Found Examined as a function of , and . , and

and the relation between and as a function of

Muddy Points What exactly is the specific impulse, How is found for rockets in space where rather than

, a measure of? (MP 11.1) ? (MP 11.2)

Why does industry use (MP 11.3)

? Is there an advantage to this?

Why isn't mechanical efficiency an issue with ramjets? (MP 11.4) How is thrust created in a ramjet? (MP 11.5)

What is the relation between of ? (MP 11.6)

and the existence of the maximum value

Why didn't we have a 2s point for the Brayton cycle with non-ideal components? (MP 11.7)

What is the variable

? (MP 11.8)

11.6.4 Turbojet Engine

Figure 11.15: Schematic with appropriate component notations added.

We now examine an engine with turbomachinery, as shown in Figure 11.15. Methodology:

1. Find thrust by finding in terms of , temperature ratios, etc. 2. Use a power balance to relate turbine parameters to compressor parameters 3. Use an energy balance across the combustor to relate the combustor temperature rise to the fuel flow rate and fuel energy content. First write-out the expressions for thrust and Isp:

where

is the fuel/air mass flow ratio,

and

Now we have to do a little algebra to manipulate these expressions into more useful forms. First we write an expression for the exit velocity:

Noting that

we can write

Thus (11..1)

which expresses the exit temperature as a function of the inlet temperature, the Mach number, and the temperature changes across each component. We now write the pressure at the exit in a similar manner:

Since

and

we write

and then equate this to our expression for the temperature in (11.1):

and

We now continue on the path to our expression for

Therefore

Now we have two steps left. First we write in terms of , by noting that they are related by the condition that the power used by the compressor is equal to the power extracted by the turbine. Second, we put the burner temperature ratio in terms of the exit temperature of the burner, ( or more specifically ) since this is the hottest point in the engine and is a frequent benchmark used for judging various designs.

The steady flow energy equation states

Assuming that the compressor and turbine are adiabatic, then

Since the turbine shaft is connected to the compressor shaft

Assuming the mass flow and specific heats are the same between compressor and turbine, this can be rewritten as

where

so

or

That was the first step relating the temperature rise across the turbine to that across the compressor. The remaining step is to write the temperature rise across the combustor in terms of .

and for an engine with an afterburner

Now substituting our expressions for , and the first expression we wrote for thrust, we get:

into our expression for

, and finally into

This is what we were seeking, an expression for thrust in terms of important design parameters and flight parameters:

By adding and substracting the quantity

we may write this write in another form which is often used,

A recap of the important variables:

Our final step involves writing the specific impulse and other measures of efficiency in terms of these same parameters. We begin by writing the First Law across the combustor to relate the fuel flow rate and heating value of the fuel to the total enthalpy rise.

and

where again,

is the fuel/air mass flow ratio. The specific impulse becomes

where is expressed in terms of typical design parameters, flight conditions, and physical constants

Similarly, we can write our overall efficiency as

or

Using the definition from before, the ideal thermal efficiency is

and the propulsive efficiency can be found as

We can now use these equations to better understand the performance of a simple turbojet engine. We will use the following parameters (with ):

Table 11.1: Turbojet Mission Parameters Mach number Altitude Ambient Temp. Speed of sound 0 0.85 2.0 Sea level 12 km 18 km 288 K 217 K 217 K 340 m/s 295 m/s 295 m/s

Note it is more typical to work with the compressor pressure ratio ( temperature ratio (

) rather than the

) so we will substitute the isentropic relationship:

into the equations before plotting the results in Figures 11.16, 11.17, and 11.18.

Figure 11.16: Performance of an ideal turbojet engine as a function of flight Mach number.

Figure 11.17: Performance of an ideal turbojet engine as a function of compressor pressure ratio and flight Mach number.

Figure 11.18: Performance of an ideal turbojet engine as a function of compressor pressure ratio and turbine inlet temperature.

11.6.5 Effect of Departures from Ideal Behavior -- Real Cycle behavior


To conclude this chapter, we will now improve our estimates of cycle performance by including the effects of irreversibility. We will use the Brayton cycle as an example. What are the sources of non-ideal performance and departures from reversibility?

Losses (entropy production) in the compressor and the turbine. Stagnation pressure decrease in the combustor. Heat transfer.

We take into account here only irreversibility in the compressor and in the turbine.

Because of these irreversibilities, we need more work, (the changes in kinetic energy from inlet to exit of the compressor are neglected), to drive the compressor than in the ideal situation. We also get less work, , back from the turbine. The consequence, as can be inferred from Figure 11.19, is that the net work from the engine is less than in the cycle with ideal components.

Figure 11.19: Gas turbine engine (Brayton) cycle showing effect of departure from ideal behavior in compressor and turbine

To develop a quantitative description of the effect of these departures from reversible behavior, consider a perfect gas with constant specific heats and neglect kinetic energy at the inlet and exit of the turbine and compressor. We define the turbine adiabatic efficiency as

where is specified to be at the same pressure ratio as for the compressor, the compressor adiabatic efficiency:

. There is a similar metric

again for the same pressure ratio. Note that for the turbine the ratio is the actual work delivered divided by the ideal work, whereas for the compressor the ratio is the ideal work needed divided by the actual work required. These are not thermal efficiencies, but rather measures of the degree to which the compression and expansion approach the ideal processes.

We now wish to find the net work done in the cycle and the efficiency. The net work is given either by the difference between the heat received and rejected or the work of the compressor and turbine, where the convention is that heat received is positive and heat rejected is negative and work done is positive and work absorbed is negative.

The thermal efficiency is

We need to calculate

From the definition of

With

Similarly, by the definition

we can find

The thermal efficiency can now be found:

With

and

or

There are several non-dimensional parameters that appear in this expression for thermal efficiency. We list these in the two sections below and show their effects in accompanying figures.

11.6.5.1 Parameters reflecting design choices

11.6.5.2 Parameters reflecting the ability to design and execute efficient components

In addition to efficiency, net rate of work is a quantity we need to examine,

Putting this in a non-dimensional form,

[Non-dimensional work as a function of cycle pressure ratio for different values of turbine entry

temperature divided by compressor entry temperature] [Overall cycle efficiency as a function of pressure ratio for different values of turbine entry temperature

divided by compressor entry temperature] [Overall cycle efficiency as a function of cycle pressure ratio for different component

efficiencies] Figure 11.20: Non-dimensional power and efficiency for a non-ideal gas turbine engine [from Cumpsty, Jet Propulsion]

Trends in net power and efficiency are shown in Figure 11.20 for parameters typical of advanced civil engines. Some points to note in the figure:

For any , the optimum pressure ratio for maximum is not the highest that can be achieved, as it is for the ideal Brayton cycle. The ideal analysis is too idealized in this regard. The highest efficiency also occurs closer to the pressure ratio for maximum power than in the case of an ideal cycle. Choosing this as a design criterion will therefore not lead to the efficiency penalty inferred from ideal cycle analysis. There is a strong sensitivity to the component efficiencies. For example, for , the cycle efficiency is roughly two-thirds of the ideal value.

The maximum power occurs at a value of max

or pressure ratio

less than that for

(this trend is captured by ideal analysis). are strongly dependent on the maximum

The maximum power and maximum temperature, .

A scale diagram of a Brayton cycle with non-ideal compressor and turbine behaviors, in terms of temperature-entropy ( - ) and pressure-volume ( - ) coordinates is given below as Figure 11.21.

[]

[]

Figure 11.21: Scale diagram of non-ideal gas turbine cycle. Nomenclature is shown in the figure. Pressure ratio , , , compressor and turbine efficiencies Cumpsty, Jet Propulsion] [from

Muddy Points Isn't it possible for the mixing of two gases to go from the final state to the initial state? If you have two gases in a box, they should eventually separate by density, right? (MP 11.9)

How can

be the maximum turbine inlet temperature? (MP 11.10)

When there are losses in the turbine that shift the expansion in - diagram to the right, does this mean there is more work than ideal since the area is greater? (MP 11.11) For an afterburning engine, why must the nozzle throat area increase if the temperature of the fluid is increased? (MP 11.12) Why doesn't the pressure in the afterburner go up if heat is added? (MP 11.13)

Why is the flow in the nozzle choked? (MP 11.14) What's the point of having a throat if it creates a retarding force? (MP 11.15) Why isn't the stagnation temperature conserved in this steady flow? (MP 11.16)

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index

UnifiedTP UnifiedTP

Thermodynamics and Propulsion

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel Cycle Contents Index Subsections

3.7.1 Work and Efficiency 3.7.2 Gas Turbine Technology and Thermodynamics 3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet 3.7.4 MIT Cogenerator

3.7 Brayton Cycle


[VW, S & B: 9.8-9.9, 9.12]

The Brayton cycle (or Joule cycle) represents the operation of a gas turbine engine. The cycle consists of four processes, as shown in Figure 3.13 alongside a sketch of an engine:

a - b Adiabatic, quasi-static (or reversible) compression in the inlet and compressor; b - c Constant pressure fuel combustion (idealized as constant pressure heat addition); c - d Adiabatic, quasi-static (or reversible) expansion in the turbine and exhaust nozzle, with which we 1. take some work out of the air and use it to drive the compressor, and 2. take the remaining work out and use it to accelerate fluid for jet propulsion, or to turn a generator for electrical power generation;

d - a Cool the air at constant pressure back to its initial condition.

Figure 3.13: Sketch of the jet engine components and corresponding thermodynamic states

The components of a Brayton cycle device for jet propulsion are shown in Figure 3.14. We will typically represent these components schematically, as in Figure 3.15. In practice, real Brayton cycles take one of two forms. Figure 3.16(a) shows an ``open'' cycle, where the working fluid enters and then exits the device. This is the way a jet propulsion cycle works. Figure 3.16(b) shows the alternative, a closed cycle, which recirculates the working fluid. Closed cycles are used, for example, in space power generation.

Figure 3.14: Schematics of typical military gas turbine engines. Top: turbojet with afterburning, bottom: GE F404 low bypass ratio turbofan with afterburning (Hill and Peterson, 1992).

Figure 3.15: Thermodynamic model of gas turbine engine cycle for power generation

[Open cycle operation]

[Closed cycle

operation] Figure 3.16: Options for operating Brayton cycle gas turbine engines

Muddy Points Would it be practical to run a Brayton cycle in reverse and use it as refrigerator? (MP 3.10)

3.7.1 Work and Efficiency


The objective now is to find the work done, the heat absorbed, and the thermal efficiency of

the cycle. Tracing the path shown around the cycle from first law gives (writing the equation in terms of a unit mass),

and back to

, the

Here is zero because is a function of state, and any cycle returns the system to its starting 3.2 state . The net work done is therefore

where , are defined as heat received by the system ( evaluate the heat transferred in processes - and - .

is negative). We thus need to

For a constant pressure, quasi-static process the heat exchange per unit mass is

We can see this by writing the first law in terms of enthalpy (see Section 2.3.4) or by remembering the definition of .

The heat exchange can be expressed in terms of enthalpy differences between the relevant states. Treating the working fluid as a perfect gas with constant specific heats, for the heat addition from the combustor,

The heat rejected is, similarly,

The net work per unit mass is given by

The thermal efficiency of the Brayton cycle can now be expressed in terms of the temperatures:
(3..8 )

To proceed further, we need to examine the relationships between the different

temperatures. We know that points points and , and and reversible, so ;

and

are on a constant pressure process as are . The other two legs of the cycle are adiabatic

Therefore , or, finally, . Using this relation in the expression for thermal efficiency, Eq. (3.8) yields an expression for the thermal efficiency of a Brayton cycle:

(3..9)

The temperature ratio across the compressor, . In terms of compressor temperature ratio, and using the relation for an adiabatic reversible process we can write the efficiency in terms of the compressor (and cycle) pressure ratio, which is the parameter commonly used:

(3..10)

Figure 3.17: Gas turbine engine pressures and temperatures

Figure 3.17 shows pressures and temperatures through a gas turbine engine (the PW4000, which powers the 747 and the 767).

Figure 3.18: Gas turbine engine pressure ratio trends (Janes Aeroengines, 1998)

Figure 3.19: Trend of Brayton cycle thermal efficiency with compressor pressure ratio

Equation (3.10) says that for a high cycle efficiency, the pressure ratio of the cycle should be increased. This trend is plotted in Figure 3.19. Figure 3.18 shows the history of aircraft engine pressure ratio versus entry into service, and it can be seen that there has been a large increase in cycle pressure ratio. The thermodynamic concepts apply to the behavior of real aerospace devices!

Muddy Points When flow is accelerated in a nozzle, doesn't that reduce the internal energy of the flow and therefore the enthalpy? (MP 3.11) Why do we say the combustion in a gas turbine engine is constant pressure? (MP 3.12) Why is the Brayton cycle less efficient than the Carnot cycle? (MP 3.13) If the gas undergoes constant pressure cooling in the exhaust outside the engine, is that still within the system boundary? (MP 3.14) Does it matter what labels we put on the corners of the cycle or not? (MP 3.15) Is the work done in the compressor always equal to the work done in the turbine plus work out (for a Brayton cyle)? (MP 3.16)

3.7.2 Gas Turbine Technology and Thermodynamics


The turbine entry temperature, , is fixed by materials technology and cost. (If the temperature is too high, the blades fail.) Figures 3.20 and 3.21 show the progression of the turbine entry temperatures in aeroengines. Figure 3.20 is from Rolls Royce and Figure 3.21 is from Pratt & Whitney. Note the relation between the gas temperature coming into the turbine blades and the blade melting temperature.

Figure 3.20: Rolls-Royce high temperature technology

Figure 3.21: Turbine blade cooling technology [Pratt & Whitney]

For a given level of turbine technology (in other words given maximum temperature) a design question is what should the compressor be? What criterion should be used to decide this? Maximum thermal efficiency? Maximum work? We examine this issue below.

Figure 3.22: Efficiency and work of two Brayton cycle engines

The problem is posed in Figure 3.22, which shows two Brayton cycles. For maximum

efficiency we would like as high as possible. This means that the compressor exit temperature approaches the turbine entry temperature. The net work will be less than the heat received; as the heat received approaches zero and so does the net work.

The net work in the cycle can also be expressed as

, evaluated in traversing the

cycle. This is the area enclosed by the curves, which is seen to approach zero as . The conclusion from either of these arguments is that a cycle designed for maximum thermal efficiency is not very useful in that the work (power) we get out of it is zero. A more useful criterion is that of maximum work per unit mass (maximum power per unit mass flow). This leads to compact propulsion devices. The work per unit mass is given by:

where

is the maximum turbine inlet temperature (a design constraint) and

is atmospheric

temperature. The design variable is the compressor exit temperature, as this is varied, we differentiate the expression for work with respect to

, and to find the maximum :

The first and the fourth terms on the right hand side of the above equation are both zero (the turbine entry temperature is fixed, as is the atmospheric temperature). The maximum work occurs where the derivative of work with respect to is zero:

(3..11)

To use Eq. (3.11), we need to relate

and

. We know that

Hence,

Plugging this expression for the derivative into Eq. (3.11) gives the compressor exit temperature for maximum work as . In terms of temperature ratio,

The condition for maximum work in a Brayton cycle is different than that for maximum efficiency. The role of the temperature ratio can be seen if we examine the work per unit mass which is delivered at this condition:

Ratioing all temperatures to the engine inlet temperature,

To find the power the engine can produce, we need to multiply the work per unit mass by the mass flow rate:
(3..12 )

The trend of work output vs. compressor pressure ratio, for different temperature ratios , is shown in Figure 3.23.

Figure 3.23: Trend of cycle work with compressor pressure ratio, for different temperature ratios

[Gas turbine engine core]

[Core power vs. turbine entry

temperature] Figure 3.24: Aeroengine core power [Koff/Meese, 1995]

Figure 3.24 shows the expression for power of an ideal cycle compared with data from

actual jet engines. Figure 3.24(a) shows the gas turbine engine layout including the core (compressor, burner, and turbine). Figure 3.24(b) shows the core power for a number of different engines as a function of the turbine rotor entry temperature. The equation in the figure for horsepower (HP) is the same as that which we just derived, except for the conversion factors. The analysis not only shows the qualitative trend very well but captures much of the quantitative behavior too. A final comment (for this section) on Brayton cycles concerns the value of the thermal efficiency. The Brayton cycle thermal efficiency contains the ratio of the compressor exit temperature to atmospheric temperature, so that the ratio is not based on the highest temperature in the cycle, as the Carnot efficiency is. For a given maximum cycle temperature, the Brayton cycle is therefore less efficient than a Carnot cycle.

Muddy Points

What are the units of

in

? (MP 3.17)

Question about the assumptions made in the Brayton cycle for maximum efficiency and maximum work (MP 3.18) You said that for a gas turbine engine modeled as a Brayton cycle the work done is , where is the heat added and is the heat rejected. Does this suggest that the work that you get out of the engine doesn't depend on how good your compressor and turbine are? since the compression and expansion were modeled as adiabatic. (MP 3.19)

3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet


A schematic of a ramjet is given in Figure 3.25.

Figure 3.25: Ideal ramjet [J. L. Kerrebrock, Aircraft Engines and Gas Turbines]

In the ramjet there are ``no moving parts.'' The processes that occur in this propulsion

device are:

: Isentropic diffusion (slowing down) and compression, with a decrease in Mach number, . : Constant pressure combustion. : Isentropic expansion through the nozzle.

The ramjet thermodynamic cycle efficiency can be written in terms of flight Mach number, , as follows:

and

so

See also Section 11.6.3 for other figures of merit.

Muddy Points Why don't we like the numbers 1 and 2 for the stations? Why do we go 0-3? (MP 3.20) For the Brayton cycle efficiency, why does ? (MP 3.21)

3.7.4 MIT Cogenerator


MIT operates a Brayton cycle power generator on campus. For more information, see the website at https://cogen.mit.edu/ctg.cfm .

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel Cycle Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2 Generalized Representation of Contents Index

3.3 The Carnot Cycle


A Carnot cycle is shown in Figure 3.4. It has four processes. There are two adiabatic reversible legs and two isothermal reversible legs. We can construct a Carnot cycle with many different systems, but the concepts can be shown using a familiar working fluid, the ideal gas. The system can be regarded as a chamber enclosed by a piston and filled with this ideal gas.

Figure 3.4: Carnot cycle -- thermodynamic diagram on left and schematic of the different stages in the cycle for a system composed of an ideal gas on the right

The four processes in the Carnot cycle are:


1. The system is at temperature at state . It is brought in contact with a heat reservoir, which is just a liquid or solid mass of large enough extent such that its temperature does not change appreciably when some amount of heat is transferred to the system. In other words, the heat reservoir is a constant temperature source (or receiver) of heat. The system then undergoes an isothermal expansion from to , with heat absorbed . 2. At state , the system is thermally insulated (removed from contact with the heat reservoir) and then let expand to . During this expansion the temperature decreases to ) . The

heat exchanged during this part of the cycle,

3. At state

the system is brought in contact with a heat reservoir at temperature

. It is

then compressed to state , rejecting heat in the process. 4. Finally, the system is compressed adiabatically back to the initial state exchange .

. The heat

The thermal efficiency of the cycle is given by the definition

(3..4)

In this equation, there is a sign convention implied. The quantities

as defined are

the magnitudes of the heat absorbed and rejected. The quantities , on the other hand are defined with reference to heat received by the system. In this example, the former is negative and the latter is positive. The heat absorbed and rejected by the system takes place during isothermal processes and we already know what their values are from Eq. (3.1):

The efficiency can now be written in terms of the volumes at the different states as (3..5)

The path from states to and from to reversible adiabatic process we know that of state, we have Along the curve ,

are both adiabatic and reversible. For a . Using the ideal gas equation , therefore, .

. Along curve . Thus,

Comparing the expression for thermal efficiency Eq. (3.4) with Eq. (3.5) shows two consequences. First, the heats received and rejected are related to the temperatures of the isothermal parts of the cycle by

(3..6)

Second, the efficiency of a Carnot cycle is given compactly by

(3..7)

The efficiency can be 100% only if the temperature at which the heat is rejected is zero. The heat and work transfers to and from the system are shown schematically in Figure 3.5.

Figure 3.5: Work and heat transfers in a Carnot cycle between two heat reservoirs

Muddy Points

Since the ,

, looking at the

graph, does that mean the farther apart

isotherms are, the greater efficiency? And that if they were very close, it would

be very inefficient? (MP 3.2) In the Carnot cycle, why are we only dealing with volume changes and not pressure changes on the adiabats and isotherms? (MP 3.3) Is there a physical application for the Carnot cycle? Can we design a Carnot engine for a propulsion device? (MP 3.4) How do we know which cycles to use as models for real processes? (MP 3.5)

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2 Generalized Representation of Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 11.2 Thermal and Propulsive Up: 11. Aircraft Engine Performance Previous: 11. Aircraft Engine Performance Contents Index

11.1 Overall Efficiency

Thus

Next: 11.2 Thermal and Propulsive Up: 11. Aircraft Engine Performance Previous: 11. Aircraft Engine Performance Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 13.4 Aircraft Endurance Up: 13. Aircraft Performance Previous: 13.2 Power Required Contents Index Subsections

13.3.1 Relation of overall efficiency, , and thermal efficiency 13.3.2 The Propulsion Energy Conversion Chain

13.3 Aircraft Range: the Breguet Range Equation


Consider an aircraft in steady, level flight, with weight , as shown in Figure 13.1. The rate of change of the gross weight of the vehicle is equal to the fuel weight flow:

For steady, level flight,

, or

The rate of change of aircraft gross weight is thus

Suppose

and

remain constant along the flight path:

We can integrate this equation for the change in aircraft weight to yield a relation between the weight change and the time of flight:

where

is the initial weight. If

is the final weight of vehicle and , is

, the relation

between vehicle parameters and flight time,

The range is the flight time multiplied by the flight speed, or,

The above equation is known as the Breguet range equation. It shows the influence of aircraft, propulsion system, and structural design parameters.

13.3.1 Relation of overall efficiency, efficiency


Suppose

, and thermal

is the heating value (``heat of combustion'') of the fuel (i.e., the energy per unit of , so

fuel mass), in J/kg. The rate of energy release is

and

Thus

and

or

Keep in mind that, in general,

13.3.2 The Propulsion Energy Conversion Chain


The above concepts are depicted in Figure 13.4 as parts of the propulsion energy conversion train mentioned in Part I, which shows the process from chemical energy contained in the fuel to energy useful to the vehicle.

Figure 13.4: The propulsion energy conversion chain from Part I

The combustion efficiency is near unity unless conditions are far off design. We can therefore regard the two main drivers as the thermal and propulsive 13.1 efficiencies. The evolution of the overall efficiency of aircraft engines in terms of these quantities was shown in Figure 11.8.

Next: 13.4 Aircraft Endurance Up: 13. Aircraft Performance Previous: 13.2 Power Required Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.5 The Internal combustion Up: 3. The First Law Previous: 3.3 The Carnot Cycle Contents Index Subsections
o

3.4.0.1 Refrigerator Hardware

3.4 Refrigerators and Heat Pumps


The Carnot cycle has been used for power, but we can also run it in reverse. If so, there is now net work into the system and net heat out of the system. There will be a quantity of heat rejected at the higher temperature and a quantity of heat absorbed at the lower temperature. The former of these is negative according to our convention and the latter is positive. The result is that work is done on the system, heat is extracted from a low temperature source and rejected to a high temperature source. The words ``low'' and ``high'' are relative and the low temperature source might be a crowded classroom on a hot day, with the heat extraction being used to cool the room. The cycle and the heat and work transfers are indicated in Figure 3.6. In this mode of operation the cycle works as a refrigerator or heat pump. ``What we pay for'' is the work, and ``what we get'' is the amount of heat extracted. A metric for devices of this type is the coefficient of performance, defined as

Figure 3.6: Operation of a Carnot refrigerator

For a Carnot cycle we know the ratios of heat in to heat out when the cycle is run forward and, since the cycle is reversible, these ratios are the same when the cycle is run in reverse. The coefficient of performance is thus given in terms of the absolute temperatures as

This can be much larger than unity.

The Carnot cycles that have been drawn are based on ideal gas behavior. For different working media, however, they will look different. We will see an example when we discuss two-phase situations. What is the same whatever the medium is the efficiency for all Carnot cycles operating between the same two temperatures.

3.4.0.1 Refrigerator Hardware


Typically the thermodynamic system in a refrigerator analysis will be a working fluid, a refrigerant, that circulates around a loop, as shown in Figure 3.7. The internal energy (and temperature) of the refrigerant is alternately raised and lowered by the devices in the loop. The working fluid is colder than the refrigerator air at one point and hotter than the surroundings at another point. Thus heat will flow in the appropriate direction, as shown by the two arrows in the heat exchangers.

Figure 3.7: Schematic of a domestic refrigerator

Starting in the upper right hand corner of the diagram, we describe the process in more detail. First the refrigerant passes through a small turbine or through an expansion valve. In these devices, work is done by the refrigerant so its internal energy is lowered to a point where the temperature of the refrigerant is lower than that of the air in the refrigerator. A heat exchanger is used to transfer energy from the inside of the refrigerator to the cold refrigerant. This lowers the internal energy of the inside and raises the internal energy of the refrigerant. Then a pump or compressor is used to do work on the refrigerant, adding additional energy to it and thus further raising its internal energy. Electrical energy is used to drive the pump or compressor. The internal energy of the refrigerant is raised to a point where its temperature is hotter than the temperature of the surroundings. The refrigerant is then passed through a heat exchanger (often coils at the back of the refrigerator) so that energy is transferred from the refrigerant to the surroundings. As a result, the internal energy of the refrigerant is reduced and the internal energy of the surroundings is increased. It is at this point where the internal energy of the contents of the refrigerator and the energy used to drive the compressor or pump are transferred to the surroundings. The refrigerant then continues on to the turbine or expansion valve, repeating the cycle.

Next: 3.5 The Internal combustion Up: 3. The First Law Previous: 3.3 The Carnot Cycle Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The First Law Contents Index

3.1 Some Properties of Engineering Cycles; Work and Efficiency


As preparation for our discussion of cycles (and as a foreshadowing of the second law), we examine two types of processes that concern interactions between heat and work. The first of these represents the conversion of work into heat. The second, which is much more useful, concerns the conversion of heat into work. The question we will pose is how efficient can this conversion be in the two cases.

Figure 3.1: Examples of the conversion of work into heat

Three examples of the first process are given in Figure 3.1. The first is the pulling of a block on a rough horizontal surface by a force which moves through some distance. Friction resists the pulling. After the force has moved through the distance, it is removed. The block then has no kinetic energy and the same potential energy it had when the force started to act. If we measured the temperature of the block and the surface we would find that it was higher than when we started. (High temperatures can be reached if the velocities of pulling are high; this is the basis of inertia welding.) The work done to move the block has been converted totally to heat. The second example concerns the stirring of a viscous liquid. There is work associated with the torque exerted on the shaft turning through an angle. When the stirring stops, the fluid comes to rest and there is (again) no change in kinetic or potential energy from the initial state. The fluid and the paddle wheels will be found to be hotter than when we started, however. The final example is the passage of a current through a resistance. This is a case of electrical work being converted to heat, indeed it models operation of an electrical heater. All the examples in Figure 3.1 have 100% conversion of work into heat. This 100% conversion could go on without limit as long as work were supplied. Is this true for the conversion of heat into work? To answer the last question, we need to have some basis for judging whether work is done in a given process. One way to do this is to ask whether we can construct a way that the process could result in the raising of a weight in a gravitational field. If so, we can say ``Work has been done.'' It may sometimes be difficult to make the link between a complicated thermodynamic process and the simple raising of a weight, but this is a rigorous test for the existence of work. One example of a process in which heat is converted to work is the isothermal (constant temperature) expansion of an ideal gas, as sketched in Figure 3.2. The system is the gas inside the chamber. As the gas expands, the piston does work on some external device. For an ideal gas, the internal energy is a function of temperature only, so that if the temperature is constant for some process the internal energy change is zero. To keep the temperature constant during the expansion, heat must be supplied. Because first law takes the form work. the

. This is a process that has 100% conversion of heat into

Figure 3.2: Isothermal expansion

The work exerted by the system is given by

where 1 and 2 denote the two states at the beginning and end of the process. The equation of state for an ideal gas is

with the number of moles of the gas contained in the chamber. Using the equation of state, the expression for work can be written as

(3..1)

For an isothermal process, , so that written in terms of the pressures at the beginning and end as

. The work can be

(3..2)

The lowest pressure to which we can expand and still receive work from the system is atmospheric pressure. Below this, we would have to do work on the system to pull the piston out further. There is thus a bound on the amount of work that can be obtained in the isothermal expansion; we cannot continue indefinitely. For a power or propulsion system, however, we would like a source of continuous power, in other words a device that would give power or propulsion as long as fuel was added to it. To do this, we need a series of processes where the system does not progress through a one-way transition from an initial state to a different final state, but rather cycles back to the initial state. What is looked for is in fact a thermodynamic cycle for the system.

We define several quantities for a cycle:


is the heat absorbed by the system. is the heat rejected by the system. is the net work done by the system.

The cycle returns to its initial state, so the overall energy change, , is zero. The net work done by the system is related to the magnitudes of the heat absorbed and the heat rejected by

The thermal efficiency of the cycle is the ratio of the work done to the heat absorbed. (Efficiencies are often usefully portrayed as ``What you get'' versus ``What you pay for.'' Here what we get is work and what we pay for is heat, or rather the fuel that generates the heat.) In terms of the heat absorbed and rejected, the thermal efficiency is

(3..3)

The thermal efficiency can only be 100% (complete conversion of heat into work) if ; a basic question is what is the maximum thermal efficiency for any arbitrary cycle? We examine this for several cases, including the Carnot cycle and the Brayton (or Joule) cycle, which is a model for the power cycle in a jet engine.

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

2.1 First Law of Thermodynamics


[VW, S & B: 2.6]

Observation leads to the following two assertions:


1. There exists for every system a property called energy, . The system energy can be considered as a sum of internal energy, kinetic energy, potential energy, and chemical energy. 1. Like the Zeroth Law, which defined a useful property, ``temperature,'' the First Law defines a useful property called ``energy.'' 2. The two new terms (compared to what you have seen in physics and dynamics, for example) are the internal energy and the chemical energy. For most situations in this class, we will neglect the chemical energy. We will generally not, however, neglect the internal energy, . It arises from the random or disorganized motion of molecules in the system, as shown in Figure 2.1. Since this molecular motion is primarily a function of temperature, the internal energy is sometimes called ``thermal energy.''

Figure 2.1: Random motion is the physical basis for internal energy

3. The internal energy,

, is a function of the state of the system.

Thus , or , or . Recall that for pure substances the entire state of the system is specified if any two properties are specified. (We will discuss the equations that relate the internal energy to these other variables as the course progresses.)
2. The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system, (2..1)

3. 4. where is the energy of the system,

is the heat input to the system, and is the work done by the system.

(thermal energy) + 1. Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5]. 2. The equation can also be written on a per unit mass basis

3. 4. In many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then

5. 6. and

7.

8. Note that and are not functions of state, but , which arises from molecular motion (see above), depends only on the state of the system; does not depend on how the system got to that state. We therefore have the striking result that:

9.

10. Sometimes this difference is emphasized by writing the First Law in differential form,
(2..2)

11. 12. where the symbol `` '' is used to denote that these are not exact differentials but rather are dependent on path. 13. Note that the signs are important:

is defined to be positive if it is transferred to the system; thus the

numerical value we substitute for will be positive if heat is transferred to the system from the surroundings, and negative if heat is transferred from the system to the surroundings. [VW, S & B: 4.7-4.8] is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] ,

14. For quasi-static processes we can substitute

15.

To give an example of where the first law is applied, consider the device shown in Figure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) is applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law -- in one form or another.

Figure 2.2: The change in energy of a system relates the heat added to the work done

The form of the first law we have given here is sometimes called the ``control mass'' form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like a jet engine for example). We will call this the ``control volume'' form of the first law [VW, S & B: 5.8-5.12].

Muddy Points What are the conventions for work and heat in the first law? (MP 2.1)

When does

? (MP 2.2)

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

2.1 First Law of Thermodynamics


[VW, S & B: 2.6]

Observation leads to the following two assertions:


1. There exists for every system a property called energy, . The system energy can be considered as a sum of internal energy, kinetic energy, potential energy, and chemical energy. 1. Like the Zeroth Law, which defined a useful property, ``temperature,'' the First Law defines a useful property called ``energy.'' 2. The two new terms (compared to what you have seen in physics and dynamics, for example) are the internal energy and the chemical energy. For most situations in this class, we will neglect the chemical energy. We will generally not, however, neglect the internal energy, . It arises from the random or disorganized motion of molecules in the system, as shown in Figure 2.1. Since this molecular motion is primarily a function of temperature, the internal energy is sometimes called ``thermal energy.''

Figure 2.1: Random motion is the physical basis for internal energy

3. The internal energy,

, is a function of the state of the system.

Thus , or , or . Recall that for pure substances the entire state of the system is specified if any two properties are specified. (We will discuss the equations that relate the internal energy to these other variables as the course progresses.)
2. The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system, (2..1)

3. 4. where is the energy of the system,

is the heat input to the system, and is the work done by the system. (thermal energy) + 1. Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5]. 2. The equation can also be written on a per unit mass basis

3. 4. In many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then

5. 6. and

7.

8. Note that and are not functions of state, but , which arises from molecular motion (see above), depends only on the state of the system; does not depend

on how the system got to that state. We therefore have the striking result that:

9.

10. Sometimes this difference is emphasized by writing the First Law in differential form,
(2..2)

11. 12. where the symbol `` '' is used to denote that these are not exact differentials but rather are dependent on path. 13. Note that the signs are important:

is defined to be positive if it is transferred to the system; thus the numerical value we substitute for will be positive if heat is transferred to the system from the surroundings, and negative if heat is transferred from the system to the surroundings. [VW, S & B: 4.7-4.8] is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] ,

14. For quasi-static processes we can substitute

15.

To give an example of where the first law is applied, consider the device shown in Figure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) is applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law -- in one form or another.

Figure 2.2: The change in energy of a system relates the heat added to the work done

The form of the first law we have given here is sometimes called the ``control mass'' form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like a jet engine for example). We will call this the ``control volume'' form of the first law [VW, S & B: 5.8-5.12].

Muddy Points What are the conventions for work and heat in the first law? (MP 2.1) When does ? (MP 2.2)

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 3. The First Law Up: 2. The First Law Previous: 2.5 Control volume form Contents Index

2.6 Muddiest Points on Chapter 2

MP 2..1 What are the conventions for work and heat in the first law?

Heat is positive if it is given to the system. Work is positive if it is done by the system.
MP 2..2 When does ?

We deal with changes in energy. When the changes in the other types of energy (kinetic, potential, strain, etc.) can be neglected compared to the changes in thermal energy, then it is a good approximation to use as representing the total energy change.
MP 2..3 When is enthalpy the same in initial and final states?

Initial and final stagnation enthalpy is the same if the flow is steady and if there is no net shaft work plus heat transfer. If the change in kinetic energy is negligible, the initial and final enthalpy is the same. The tank problem is unsteady so the initial and final enthalpies are not the same. See the discussion of the steady flow energy equation in notes, Section 2.5.
MP 2..4 In the filling of a tank, why (physically) is the final temperature in the tank higher than the initial temperature?

Work is done on the system, which in this problem is the mass of gas that is pushed into the tank.
MP 2..5 What is shaft work?

I am not sure how best to answer, but it appears that the difficulty people are having might be associated with being able to know when one can say that shaft work occurs. There are several features of a process that produces (or absorbs) shaft work. First of all the view taken of the process is one of control volume, rather than control mass (see the discussion of control volumes in Chapter I or in IAW). Second, there needs to be a shaft or equivalent device (a moving belt, a row of blades) that can be identified as the work carrier. Third, the shaft work is work over and above the ``flow work'' that is done by (or received by) the streams that exit and enter the control volume.
MP 2..6 What distinguishes shaft work from other works?

The term shaft work arises in using a control volume approach. As we have defined it, ``shaft work'' is all work over and above work associated with the ``flow work'' (the work done by pressure forces). Generally this means work done by rotating machinery, which is carried by a shaft from the control volume to the outside world. There could also be work over and above the pressure force work done by shear stresses at the boundaries of the control volume, but this is seldom important if the control boundary is normal to the flow direction. If we consider a system (a mass of fixed identity, say a blob of gas) flowing through some device, neglecting the effects of raising or lowering the blob the only mode of work would be the work to compress the blob. This would be true even if the blob were flowing through a turbine or compressor. (In doing this we are focusing on the same material as it undergoes the unsteady compression or expansion processes in the device, rather than looking at a

control volume, through which mass passes.) The question about shaft work and non shaft work has been asked several times. I am not sure how best to answer, but it appears that the difficulty people are having might be associated with being able to know when one can say that shaft work occurs. There are several features of a process that produces (or absorbs) shaft work. First of all, the view taken of the process is one of control volume, rather than control mass (see the discussion of control volumes in Section 2.5 or in IAW). Second, there needs to be a shaft or equivalent device (a moving belt, a row of blades) that can be identified as the work carrier. Third, the shaft work is work over and above the flow work that is done by (or received by) the streams that exit and enter the control volume.
MP 2..7 Definition of a control volume.

A control volume is an enclosure that separates a quantity of matter from the surroundings or environment. The enclosure does not necessarily have to consist of a solid boundary like the walls of a vessel. It is only necessary that the enclosure forms a closed surface and that its properties are defined everywhere. An enclosure may transmit heat or be a heat insulator. It may be deformable and thus capable of transmitting work to the system. It may also be capable of transmitting mass.
MP 2..8 What is the difference between enthalpy and stagnation enthalpy?

The enthalpy of a gas is defined as , and represents both the internal energy of that state and the flow work done on the gas to get it at that pressure and density. The stagnation enthalpy of a gas is defined as both the enthalpy and the kinetic energy of the gas at that state. and accounts for

Next: 3. The First Law Up: 2. The First Law Previous: 2.5 Control volume form Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

2.1 First Law of Thermodynamics

[VW, S & B: 2.6]

Observation leads to the following two assertions:


1. There exists for every system a property called energy, . The system energy can be considered as a sum of internal energy, kinetic energy, potential energy, and chemical energy. 1. Like the Zeroth Law, which defined a useful property, ``temperature,'' the First Law defines a useful property called ``energy.'' 2. The two new terms (compared to what you have seen in physics and dynamics, for example) are the internal energy and the chemical energy. For most situations in this class, we will neglect the chemical energy. We will generally not, however, neglect the internal energy, . It arises from the random or disorganized motion of molecules in the system, as shown in Figure 2.1. Since this molecular motion is primarily a function of temperature, the internal energy is sometimes called ``thermal energy.''

Figure 2.1: Random motion is the physical basis for internal energy

3. The internal energy,

, is a function of the state of the system.

Thus , or , or . Recall that for pure substances the entire state of the system is specified if any two properties are specified. (We will discuss the equations that relate the internal energy to these other variables as the course progresses.)
2. The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system, (2..1)

3. 4. where is the energy of the system,

is the heat input to the system, and is the work done by the system. (thermal energy) + 1. Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5].

2. The equation can also be written on a per unit mass basis

3. 4. In many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then

5. 6. and

7.

8. Note that and are not functions of state, but , which arises from molecular motion (see above), depends only on the state of the system; does not depend on how the system got to that state. We therefore have the striking result that:

9.

10. Sometimes this difference is emphasized by writing the First Law in differential form,
(2..2)

11. 12. where the symbol `` '' is used to denote that these are not exact differentials but rather are dependent on path. 13. Note that the signs are important:

is defined to be positive if it is transferred to the system; thus the numerical value we substitute for will be positive if heat is transferred to the system from the surroundings, and negative if heat is transferred from the system to the surroundings. [VW, S & B: 4.7-4.8]

is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] ,

14. For quasi-static processes we can substitute

15.

To give an example of where the first law is applied, consider the device shown in Figure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) is applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law -- in one form or another.

Figure 2.2: The change in energy of a system relates the heat added to the work done

The form of the first law we have given here is sometimes called the ``control mass'' form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like a jet engine for example). We will call this the ``control volume'' form of the first law [VW, S & B: 5.8-5.12].

Muddy Points What are the conventions for work and heat in the first law? (MP 2.1) When does ? (MP 2.2)

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

2.1 First Law of Thermodynamics


[VW, S & B: 2.6]

Observation leads to the following two assertions:


1. There exists for every system a property called energy, . The system energy can be considered as a sum of internal energy, kinetic energy, potential energy, and chemical energy. 1. Like the Zeroth Law, which defined a useful property, ``temperature,'' the First Law defines a useful property called ``energy.'' 2. The two new terms (compared to what you have seen in physics and dynamics, for example) are the internal energy and the chemical energy. For most situations in this class, we will neglect the chemical energy. We will generally not, however, neglect the internal energy, . It arises from the random or disorganized motion of molecules in the system, as shown in Figure 2.1. Since this molecular motion is primarily a function of temperature, the internal energy is sometimes called ``thermal energy.''

Figure 2.1: Random motion is the physical basis for internal energy

3. The internal energy,

, is a function of the state of the system.

Thus , or , or . Recall that for pure substances the entire state of the system is specified if any two properties are

specified. (We will discuss the equations that relate the internal energy to these other variables as the course progresses.)
2. The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system, (2..1)

3. 4. where is the energy of the system,

is the heat input to the system, and is the work done by the system. (thermal energy) + 1. Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5]. 2. The equation can also be written on a per unit mass basis

3. 4. In many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then

5. 6. and

7.

8. Note that and are not functions of state, but , which arises from molecular motion (see above), depends only on the state of the system; does not depend on how the system got to that state. We therefore have the striking result that:

9.

10. Sometimes this difference is emphasized by writing the First Law in differential form,
(2..2)

11. 12. where the symbol `` '' is used to denote that these are not exact differentials but rather are dependent on path. 13. Note that the signs are important:

is defined to be positive if it is transferred to the system; thus the numerical value we substitute for will be positive if heat is transferred to the system from the surroundings, and negative if heat is transferred from the system to the surroundings. [VW, S & B: 4.7-4.8] is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] ,

14. For quasi-static processes we can substitute

15.

To give an example of where the first law is applied, consider the device shown in Figure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) is applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law -- in one form or another.

Figure 2.2: The change in energy of a system relates the heat added to the work done

The form of the first law we have given here is sometimes called the ``control mass'' form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like a jet engine for example). We will call this the ``control volume'' form of the first law [VW, S & B: 5.8-5.12].

Muddy Points What are the conventions for work and heat in the first law? (MP 2.1) When does ? (MP 2.2)

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index Subsections

1.3.1 Heat 1.3.2 Zeroth Law of Thermodynamics 1.3.3 Work o 1.3.3.1 Example: Work on Two Simple Paths o 1.3.3.2 Example: Work Done During Expansion of a Gas

1.3.4 Work vs. Heat - which is which?

1.3 Changing the State of a System with Heat and Work


Changes in the state of a system are produced by interactions with the environment through heat and work, which are two different modes of energy transfer. During these interactions, equilibrium (a static or quasi-static process) is necessary for the equations that relate system properties to one-another to be valid.

1.3.1 Heat
Heat is energy transferred due to temperature differences only. 1. 2. 3. 4. Heat transfer can alter system states; Bodies don't ``contain'' heat; heat is identified as it comes across system boundaries; The amount of heat needed to go from one state to another is path dependent; Adiabatic processes are ones in which no heat is transferred.

1.3.2 Zeroth Law of Thermodynamics


With the material we have discussed so far, we are now in a position to describe the Zeroth Law. Like the other laws of thermodynamics we will see, the Zeroth Law is based on observation. We start with two such observations:
1. If two bodies are in contact through a thermally-conducting boundary for a sufficiently long time, no further observable changes take place; thermal equilibrium is said to prevail. 2. Two systems which are individually in thermal equilibrium with a third are in thermal equilibrium with each other; all three systems have the same value of the property called temperature.

These closely connected ideas of temperature and thermal equilibrium are expressed formally in the ``Zeroth Law of Thermodynamics:''
Zeroth Law: There exists for every thermodynamic system in equilibrium a property called temperature. Equality of temperature is a necessary and sufficient condition for thermal equilibrium.

The Zeroth Law thus defines a property (temperature) and describes its behavior1.3. Note that this law is true regardless of how we measure the property temperature. (Other relationships we work with will typically require an absolute scale, so in these notes we use either the Kelvin or Rankine scales. Temperature scales will be discussed further in Section 6.2.) The zeroth law is depicted schematically in Figure 1.8.

Figure 1.8: The zeroth law schematically

1.3.3 Work
[VW, S & B: 4.1-4.6] Section 1.3.1 stated that heat is a way of changing the energy of a system by virtue of a temperature difference only. Any other means for changing the energy of a system is called work. We can have push-pull work (e.g. in a piston-cylinder, lifting a weight), electric and magnetic work (e.g. an electric motor), chemical work, surface tension work, elastic work, etc. In defining work, we focus on the effects that the system (e.g. an engine) has on its surroundings. Thus we define work as being positive when the system does work on the surroundings (energy leaves the system). If work is done on the system (energy added to the system), the work is negative. Consider a simple compressible substance, for example, a gas (the system), exerting a force on the surroundings via a piston, which moves through some distance, (Figure 1.9). The work doneon the surroundings, , is

therefore

Why is the pressure

instead of

? Consider

(vacuum). No work is done on the

surroundings even though

changes and the system volume changes.

Use of instead of is often inconvenient because it is usually the state of the system that we are interested in. The external pressure can only be related to the system pressure if . For this to occur, there cannot be any friction, and the process must also be slow enough so that pressure differences due to accelerations are not significant. In other words, we require a ``quasi-static'' process, . Consider .

Therefore, when

is small (the process is quasi-static),

and the work done by the system is the same as the work done on the surroundings.

Under these conditions, we say that the process is ``reversible.'' The conditions for reversibility are that:
1. If the process is reversed, the system and the surroundings will be returned to the original states. 2. To reverse the process we need to apply only an infinitesimal . A reversible process can be altered in direction by infinitesimal changes in the external conditions (see Van Ness, Chapter 2).

Remember this result, that we can only relate work done on surroundings to system pressure for quasi-static (or reversible) processes. In the case of a ``free expansion,'' where (vacuum), is not related to because the system is not in equilibrium. (and thus, not related to the work)

We can write the above expression for work done by the system in terms of the specific volume, ,

where is the mass of the system. Note that if the system volume expands against a force, work is done by the system. If the system volume contracts under a force, work is done on the system.

Figure 1.9: A closed system (dashed box) against a piston, which moves into the surroundings

Figure 1.10: Work during an irreversible process

For simple compressible substances in reversible processes, the work done can be represented as the area under a curve in a pressure-volume diagram, as in Figure 1.11(a).

[Work is area under curve of

[Work depends on

path] Figure 1.11: Work in coordinates

Key points to note are the following:


1. Properties only depend on states, but work is path dependent (depends on the path taken between states); therefore work is not a property, and not a state variable. 2. When we say , the work between states 1 and 2, we need to specify the path. ; either the work must be

3. For irreversible (non-reversible) processes, we cannot use given or it must be found by another method.

Muddy Points How do we know when work is done? (MP 1.3)

1.3.3.1 Example: Work on Two Simple Paths


Consider Figure 1.12, which shows a system undergoing quasi-static processes for which we can calculate work interactions as .

Figure 1.12: Simple processes

Along Path a:

Along Path b:

Practice Questions Given a piston filled with air, ice, a bunsen burner, and a stack of small weights, describe
1. how you would use these to move along either path a or path b above, and 2. how you would physically know the work is different along each path.

1.3.3.2 Example: Work Done During Expansion of a Gas


Consider the quasi-static, isothermal expansion of a thermally ideal gas from , to

, , as shown in Figure 1.13. To find the work we must know the path. Is it specified? Yes, the path is specified as isothermal.

Figure 1.13:Quasi-static, isothermal expansion of an ideal gas

The equation of state for a thermally ideal gas is

where is the number of moles, is the Universal gas constant, and is the total system volume. We write the work as above, substituting the ideal gas equation of state,

also for

, so the work done by the system is

or in terms of the specific volume and the system mass,

1.3.4 Work vs. Heat - which is which?


We can have one, the other, or both: it depends on what crosses the system boundary (and thus, on how we define our system). For example consider a resistor that is heating a volume of water (Figure 1.14):

Figure 1.14: A resistor heating water

1. If the water is the system, then the state of the system will be changed by heat transferred from the resistor. 2. If the system is the water and the resistor combined, then the state of the system will be changed by electrical work.

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index Subsections

1.3.1 Heat 1.3.2 Zeroth Law of Thermodynamics 1.3.3 Work o 1.3.3.1 Example: Work on Two Simple Paths o 1.3.3.2 Example: Work Done During Expansion of a Gas

1.3.4 Work vs. Heat - which is which?

1.3 Changing the State of a System with Heat and Work


Changes in the state of a system are produced by interactions with the environment through heat and work, which are two different modes of energy transfer. During these interactions, equilibrium (a static or quasi-static process) is necessary for the equations that relate system properties to one-another to be valid.

1.3.1 Heat
Heat is energy transferred due to temperature differences only. 1. 2. 3. 4. Heat transfer can alter system states; Bodies don't ``contain'' heat; heat is identified as it comes across system boundaries; The amount of heat needed to go from one state to another is path dependent; Adiabatic processes are ones in which no heat is transferred.

1.3.2 Zeroth Law of Thermodynamics


With the material we have discussed so far, we are now in a position to describe the Zeroth Law. Like the other laws of thermodynamics we will see, the Zeroth Law is based on

observation. We start with two such observations:


1. If two bodies are in contact through a thermally-conducting boundary for a sufficiently long time, no further observable changes take place; thermal equilibrium is said to prevail. 2. Two systems which are individually in thermal equilibrium with a third are in thermal equilibrium with each other; all three systems have the same value of the property called temperature.

These closely connected ideas of temperature and thermal equilibrium are expressed formally in the ``Zeroth Law of Thermodynamics:''
Zeroth Law: There exists for every thermodynamic system in equilibrium a property called temperature. Equality of temperature is a necessary and sufficient condition for thermal equilibrium.

The Zeroth Law thus defines a property (temperature) and describes its behavior1.3. Note that this law is true regardless of how we measure the property temperature. (Other relationships we work with will typically require an absolute scale, so in these notes we use either the Kelvin or Rankine scales. Temperature scales will be discussed further in Section 6.2.) The zeroth law is depicted schematically in Figure 1.8.

Figure 1.8: The zeroth law schematically

1.3.3 Work
[VW, S & B: 4.1-4.6] Section 1.3.1 stated that heat is a way of changing the energy of a system by virtue of a temperature difference only. Any other means for changing the energy of a system is called work. We can have push-pull work (e.g. in a piston-cylinder, lifting a weight), electric and magnetic work (e.g. an electric motor), chemical work, surface tension work, elastic work, etc. In defining work, we focus on the effects that the system (e.g. an engine) has on

its surroundings. Thus we define work as being positive when the system does work on the surroundings (energy leaves the system). If work is done on the system (energy added to the system), the work is negative. Consider a simple compressible substance, for example, a gas (the system), exerting a force on the surroundings via a piston, which moves through some distance, (Figure 1.9). The work doneon the surroundings, , is

therefore

Why is the pressure

instead of

? Consider

(vacuum). No work is done on the

surroundings even though

changes and the system volume changes.

Use of instead of is often inconvenient because it is usually the state of the system that we are interested in. The external pressure can only be related to the system pressure if . For this to occur, there cannot be any friction, and the process must also be slow enough so that pressure differences due to accelerations are not significant. In other words, we require a ``quasi-static'' process, . Consider .

Therefore, when

is small (the process is quasi-static),

and the work done by the system is the same as the work done on the surroundings.

Under these conditions, we say that the process is ``reversible.'' The conditions for reversibility are that:
1. If the process is reversed, the system and the surroundings will be returned to the original states. 2. To reverse the process we need to apply only an infinitesimal . A reversible process can be altered in direction by infinitesimal changes in the external conditions (see Van Ness, Chapter 2).

Remember this result, that we can only relate work done on surroundings to system pressure for quasi-static (or reversible) processes. In the case of a ``free expansion,'' where (vacuum), is not related to because the system is not in equilibrium. (and thus, not related to the work)

We can write the above expression for work done by the system in terms of the specific volume, ,

where is the mass of the system. Note that if the system volume expands against a force, work is done by the system. If the system volume contracts under a force, work is done on the system.

Figure 1.9: A closed system (dashed box) against a piston, which moves into the surroundings

Figure 1.10: Work during an irreversible process

For simple compressible substances in reversible processes, the work done can be represented as the area under a curve in a pressure-volume diagram, as in Figure 1.11(a).

[Work is area under curve of

[Work depends on

path] Figure 1.11: Work in coordinates

Key points to note are the following:


1. Properties only depend on states, but work is path dependent (depends on the path taken between states); therefore work is not a property, and not a state variable. 2. When we say , the work between states 1 and 2, we need to specify the path. ; either the work must be

3. For irreversible (non-reversible) processes, we cannot use given or it must be found by another method.

Muddy Points How do we know when work is done? (MP 1.3)

1.3.3.1 Example: Work on Two Simple Paths


Consider Figure 1.12, which shows a system undergoing quasi-static processes for which we can calculate work interactions as .

Figure 1.12: Simple processes

Along Path a:

Along Path b:

Practice Questions Given a piston filled with air, ice, a bunsen burner, and a stack of small weights, describe
1. how you would use these to move along either path a or path b above, and 2. how you would physically know the work is different along each path.

1.3.3.2 Example: Work Done During Expansion of a Gas


Consider the quasi-static, isothermal expansion of a thermally ideal gas from , to

, , as shown in Figure 1.13. To find the work we must know the path. Is it specified? Yes, the path is specified as isothermal.

Figure 1.13:Quasi-static, isothermal expansion of an ideal gas

The equation of state for a thermally ideal gas is

where is the number of moles, is the Universal gas constant, and is the total system volume. We write the work as above, substituting the ideal gas equation of state,

also for

, so the work done by the system is

or in terms of the specific volume and the system mass,

1.3.4 Work vs. Heat - which is which?


We can have one, the other, or both: it depends on what crosses the system boundary (and thus, on how we define our system). For example consider a resistor that is heating a volume of water (Figure 1.14):

Figure 1.14: A resistor heating water

1. If the water is the system, then the state of the system will be changed by heat transferred from the resistor. 2. If the system is the water and the resistor combined, then the state of the system will be changed by electrical work.

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.7 Examples of Lost Up: 6. Applications of the Previous: 6.5 Irreversibility, Entropy Changes, Contents Index

6.6 Entropy and Unavailable Energy (Lost Work by Another Name)


Consider a system consisting of a heat reservoir at at in surroundings (the atmosphere) . For an amount of times the

. The surroundings are equivalent to a second reservoir at

heat, , transferred from the reservoir, the maximum work we could derive is thermal efficiency of a Carnot cycle operated between these two temperatures:

(6..7)

Only part of the heat transferred can be turned into work, in other words only part of the heat energy is available to be used as work.

Suppose we transferred the same amount of heat from the reservoir directly to another reservoir at a temperature heat, . The maximum work available from the quantity of is

, beforethe transfer to the reservoir at

The maximum amount of work available after the transfer to the reservoir at

is

There is an amount of energy that could have been converted to work prior to the irreversible heat transfer process of magnitude ,

or

However,

is the entropy gain of the reservoir at

and (

) is the entropy decrease

of the reservoir at . The amount of energy, , that could have been converted to work (but now cannot be) can therefore be written in terms of entropy changes and the temperature of the surroundings as

The situation just described is a special case of an important principle concerning entropy changes, irreversibility and the loss of capability to do work. We thus now develop it in a more general fashion, considering an arbitrary system undergoing an irreversible state change, which transfers heat to the surroundings (for example the atmosphere), which can be assumed to be at constant temperature, system during the state change is surroundings is (with . The change in internal energy of the . The change in entropy of the

the heat transfer to the system)

Now consider restoring the system to the initial state by a reversible process. To do this we need to do work, , on the system and extract from the system a quantity of heat, . (We did this, for example, in ``undoing'' the free expansion process.) The change in internal energy is (with the quantities and both regarded, in this example, as positive for work done by the surroundings and heat given to the surroundings)6.2.

In this reversible process, the entropy of the surroundings is changed by

For the combined changes (the irreversible state change and the reversible state change back to the initial state), the energy change is zero because the energy is a function of state,

Thus,

For the system, the overall entropy change for the combined process is zero, because the entropy is a function of state,

The total entropy change is thus only reflected in the entropy change of the surroundings:

The surroundings can be considered a constant temperature heat reservoir and their entropy change is given by

We also know that the total entropy change, for system plus surroundings is,

The total entropy change is associated only with the irreversible process and is related to the work in the two processes by

The quantity represents the extra work required to restore the system to the original state. If the process were reversible, we would not have needed any extra work to do this. It represents a quantity of work that is now unavailable because of the irreversibility. The quantity can also be interpreted as the work that the system would have done if the original as the

process were reversible. From either of these perspectives we can identify

quantity we denoted previously as , representing lost work. The lost work in any irreversible process can therefore be related to the total entropy change (system plus surroundings) and the temperature of the surroundings by

To summarize the results of the above arguments for processes where heat can be exchanged with the surroundings at
1.

represents the difference between work we actually obtained and work that would be done during a reversible state change. It is the extra work that would be needed to restore the system to its initial state. ; ; . .

2. For a reversible process, 3. For an irreversible process,

4. an irreversible process.

is the energy that becomes unavailable for work during

Muddy Points Is path dependent? (MP 6.11) the and going from the final state back to the initial state?

Are and (MP 6.12)

Next: 6.7 Examples of Lost Up: 6. Applications of the Previous: 6.5 Irreversibility, Entropy Changes, Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.4 Specific Heats Up: 2. The First Law Previous: 2.2 Corollaries of the Contents Index Subsections

2.3.1 Adiabatic, steady, throttling of a gas 2.3.2 Quasi-Static Expansion of a Gas 2.3.3 Transient filling of a tank 2.3.4 The First Law in Terms of Enthalpy

2.3 Example Applications of the First Law to motivate the use of a property called ``enthalpy''
[VW, S & B: 5.4-5.5]

2.3.1 Adiabatic, steady, throttling of a gas (flow through a valve or other restriction)
Figure 2.5 shows the configuration of interest. We wish to know the relation between properties upstream of the valve, denoted by ``1'' and those downstream, denoted by ``2''.

Figure 2.5: Adiabatic flow through a valve, a generic throttling process

Figure 2.6: Equivalence of actual system and piston model

To analyze this situation, we can define the system (choosing the appropriate system is often a critical element in effective problem solving) as a unit mass of gas in the following two states. Initially the gas is upstream of the valve and just through the valve. In the final state the gas is downstream of the valve plus just before the valve. The figures on the left of Figure 2.6 show the actual configuration just described. In terms of the system behavior, however, we could replace the fluid external to the system by pistons which exert the same pressure that the external fluid exerts, as indicated schematically on the right side of Figure 2.6. The process is adiabatic, with changes in potential energy and kinetic energy assumed to be negligible. The first law for the system is therefore

The work done by the system is

Use of the first law leads to

In words, the initial and final states of the system have the same value of the

quantity

. For the case examined, since we are dealing with a unit mass, the . ,

initial and final states of the system have the same value of We define this quantity as the ``enthalpy,'' usually denoted by

In terms of the specific quantities, the enthalpy per unit mass is

It is a function of the state of the system. kilogram.

has units of Joules, and

has units of Joules per

The utility and physical significance of enthalpy will become clearer as we work with more flow problems. For now, you may wish to think of it as follows (Levenspiel, 1996). When you evaluate the energy of an object of volume , you have to remember that the object had to push the surroundings out of the way to make room for itself. With pressure on the

object, the work required to make a place for itself is . This is so with any object or system, and this work may not be negligible. (The force of one atmosphere pressure on one square meter is equivalent to the force of a mass of about .) Thus the total energy of a body is its internal energy plus the extra energy it is credited with by having a volume at pressure . We call this total energy the enthalpy, .

Muddy Points When is enthalpy the same in initial and final states? (MP 2.3)

2.3.2 Quasi-Static Expansion of a Gas


Consider a quasi-static process of constant pressure expansion. We can write the first law in terms of the states before and after the expansion as

and writing the work in terms of system properties,

By grouping terms we can write the heat input in terms of the enthalpy change of the system:

2.3.3 Transient filling of a tank


Another example of a flow process, this time for an unsteady flow, is the transient process of filling a tank, initially evacuated, from a surrounding atmosphere, which is at a pressure and a temperature . The configuration is shown in Figure 2.7.

Figure 2.7: A transient problem - filling of a tank from the atmosphere

At a given time, the valve at the tank inlet is opened and the outside air rushes in. The inflow stops when the pressure inside is equal to the pressure outside. The tank is insulated, so there is no heat transfer to the atmosphere. What is the final temperature of the gas in the tank? This time we take the system to be all the gas that enters the tank. The initial state has the system completely outside the tank, and the final state has the system completely inside the tank. The kinetic energy initially and in the final state is negligible, as is the change in potential energy, so the first law again takes the form

Work is done on the system, of magnitude

, where

is the initial volume of the system, so

In terms of quantities per unit mass ( system),

, where

is the mass of the

The final value of the internal energy is

For a perfect gas with constant specific heats (see the next section, Section 2.4),

The final temperature is thus roughly

hotter than the outside air!

It may be helpful to recap what we used to solve this problem. There were basically four steps:
1. 2. 3. 4. Definition of the system Use of the first law Equating the work to a `` '' term Assuming the fluid to be a perfect gas with constant specific heats.

A message that can be taken from both of these examples (as well as from a large number of other more complex situations, is that the quantity occurs naturally in problems of fluid flow. Because the combination appears so frequently, it is not only defined but also tabulated as a function of temperature and pressure for a number of working fluids.

Muddy Points In the filling of a tank, why (physically) is the final temperature in the tank higher than the

initial temperature? (MP 2.4)

2.3.4 The First Law in Terms of Enthalpy


We start with the first law in differential form and substitute static or reversible process: for by assuming a quasi-

The definition of enthalpy,

can be differentiated (applying the chain rule to the

term) to produce

Substituting the

above for the

in the First Law, we obtain

or

Next: 2.4 Specific Heats Up: 2. The First Law Previous: 2.2 Corollaries of the Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 15.4 First Law Analysis Up: 15. Generating Heat: Thermochemistry Previous: 15.2 Fuel-Air Ratio Contents Index

15.3 Enthalpy of formation


The systems we have worked with until now have been of fixed chemical composition. Because of this, we could use thermodynamic properties relative to an arbitrary base, since all comparisons could be made with respect to the chosen base. For example, the specific energy for steam. If there are no changes in composition, and only changes in properties of given substances, this is adequate. If there are changes in composition, however, we need to have a reference state so there is consistency for different substances. The convention used is that the reference state is a temperature of ( ) and a pressure of . (These are roughly room conditions.) At these reference conditions, the enthalpy of the elements (oxygen, hydrogen, nitrogen, carbon, etc.) is taken as zero. The results of a combustion process can be diagrammed as in Figure 15.1. The reactants enter at standard conditions; the combustion (reaction) takes place in the volume indicated. Downstream of the reaction zone there is an appropriate amount of heat transfer with the surroundings so that the products leave at the standard conditions. For the reaction of carbon and oxygen to produce is , the heat that has to be extracted ; this is heat that comes out of the control volume.

Figure 15.1: Constant pressure combustion

There is no shaft work done in the control volume and the first law for the control volume (SFEE) reduces to:

We can write this statement in the form (15..1)

In Eq. (15.1) the subscripts ``R'' and ``P'' on the summations refer to the reactants ( ) and products ( ) respectively. The subscripts on the mass flow rates and enthalpies refer to all of the components at inlet and at exit.

The relation in terms of mass flows can be written in molar form, which is often more convenient for reacting flow problems, by using the molecular weight, molar mass flow rate, as , and molar enthalpy, , to define the

, for any individual ith (or eth) component

The SFEE is, in these terms, (15..2)

The statements that have been made do not necessarily need to be viewed in the context of flow processes. Suppose we have one unit of and one unit of at the initial conditions and we . If so,

carry out a constant pressure reaction at ambient pressure,

since

. Combining terms,

or

In terms of the numbers of moles and the specific enthalpy this is (15..3)

The enthalpy of

, at

and

, with reference to a base where the enthalpy of . Values of the heat of

the elements is zero, is called the enthalpy of formation and denoted by formation for a number of substances are given in Table A.9 in SB&VW.

The enthalpies of the reactants and products for the formation of

are:

The enthalpy of

in any other state

is given by

These descriptions can be applied to any compound. For elements or compounds that exist in more than one state at the reference conditions (for example, carbon exists as diamond and as graphite), we also need to specify the state.

Note that there is a minus sign for the heat of formation. The heat transfer is out of the control volume and is thus negative by our convention. This means that .

Muddy Points Is the enthalpy of formation equal to the heat transfer out of the combustion during the formation reaction? (MP 15.4) Are the enthalpies of and (monoatomic hydrogen) both zero at ? (MP 15.5)

Next: 15.4 First Law Analysis Up: 15. Generating Heat: Thermochemistry Previous: 15.2 Fuel-Air Ratio Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.9 Muddiest Points on Up: 6. Applications of the Previous: 6.7 Examples of Lost Contents Index Subsections

6.8.1 Entropy 6.8.2 Reversible and Irreversible Processes 6.8.3 Examples of Reversible and Irreversible Processes

6.8 Some Overall Comments on Entropy, Reversible and Irreversible Processes

[Mainly excerpted (with some alterations) from: Engineering Thermodynamics, William C. Reynolds and Henry C. Perkins, McGraw-Hill Book Company, 1977.]

6.8.1 Entropy
1. Entropy is a thermodynamic property that measures the degree of randomization or disorder at the microscopic level. The natural state of affairs is for entropy to be produced by all processes. 2. A macroscopic feature which is associated with entropy production is a loss of ability to do useful work. Energy is degraded to a less useful form, and it is sometimes said that there is a decrease in the availability of energy. 3. Entropy is an extensive thermodynamic property. In other words, the entropy of a complex system is the sum of the entropies of its parts. 4. The notion that entropy can be produced, but never destroyed, is the second law of thermodynamics.

6.8.2 Reversible and Irreversible Processes


Processes can be classed as reversible or irreversible. The concept of a reversible process is an important one which directly relates to our ability to recognize, evaluate, and reduce irreversibilities in practical engineering processes. Consider an isolated system. The second law says that any process that would reduce the entropy of the isolated system is impossible. Suppose a process takes place within the isolated system in what we shall call the forward direction. If the change in state of the system is such that the entropy increases for the forward process, then for the backward process (that is, for the reverse change in state) the entropy would decrease. The backward process is therefore impossible, and we therefore say that the forward process is irreversible. If a process occurs, however, in which the entropy is unchanged by the forward process, then it would also be unchanged by the reverse process. Such a process could go in either direction without contradicting the second law. Processes of this latter type are called reversible. The key idea of a reversible process is that it does not produce any entropy. Entropy is produced in irreversible processes. All real processes (with the possible exception of superconducting current flows) are in some measure irreversible, though many processes can be analyzed quite adequately by assuming that they are reversible. Some processes that are clearly irreversible include: mixing of two gases, spontaneous combustion, friction, and the transfer of energy as heat from a body at high temperature to a body at low temperature. Recognition of the irreversibilities in a real process is especially important in engineering. Irreversibility, or departure from the ideal condition of reversibility, reflects an increase in the amount of disorganized energy at the expense of organized energy. The organized energy (such as that of a raised weight) is easily put to practical use; disorganized energy (such as

the random motions of the molecules in a gas) requires ``straightening out'' before it can be used effectively. Further, since we are always somewhat uncertain about the microscopic state, this straightening can never be perfect. Consequently the engineer is constantly striving to reduce irreversibilities in systems, in order to obtain better performance.

6.8.3 Examples of Reversible and Irreversible Processes


Processes that are usually idealized as reversible include:

Frictionless movement Restrained compression or expansion Energy transfer as heat due to infinitesimal temperature nonuniformity Electric current flow through a zero resistance Restrained chemical reaction Mixing of two samples of the same substance at the same state.

Processes that are irreversible include:


Movement with friction Unrestrained expansion Energy transfer as heat due to large temperature non uniformities Electric current flow through a non zero resistance Spontaneous chemical reaction Mixing of matter of different composition or state.

Next: 6.9 Muddiest Points on Up: 6. Applications of the Previous: 6.7 Examples of Lost Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 7.2 Microscopic and Macroscopic Up: 7. Entropy on the Previous: 7. Entropy on the Contents Index

7.1 Entropy Change in Mixing of Two Ideal Gases


Consider an insulated rigid container of gas separated into two halves by a heat conducting partition so the temperature of the gas in each part is the same. One side contains air, the

other side another gas, say argon, both regarded as ideal gases. The mass of gas in each side is such that the pressure is also the same. The entropy of this system is the sum of the entropies of the two parts: . Suppose the partition is taken away so the gases are free to diffuse throughout the volume. For an ideal gas, the energy is not a function of volume, and, for each gas, there is no change in temperature. (The energy of the overall system is unchanged, the two gases were at the same temperature initially, so the final temperature is the same as the initial temperature.) The entropy change of each gas is thus the same as that for a reversible isothermal expansion from the initial specific volume specific volume, is . For a mass of ideal gas, the entropy change to the final

. The entropy change of the system is

(7..1)

Equation (7.1) states that there is an entropy increase due to the increased volume that each gas is able to access.

Examining the mixing process on a molecular level gives additional insight. Suppose we were able to see the gas molecules in different colors, say the air molecules as white and the argon molecules as red. After we took the partition away, we would see white molecules start to move into the red region and, similarly, red molecules start to come into the white volume. As we watched, as the gases mixed, there would be more and more of the different color molecules in the regions that were initially all white and all red. If we moved further away so we could no longer pick out individual molecules, we would see the growth of pink regions spreading into the initially red and white areas. In the final state, we would expect a uniform pink gas to exist throughout the volume. There might be occasional small regions which were slightly more red or slightly more white, but these fluctuations would only last for a time on the order of several molecular collisions. In terms of the overall spatial distribution of the molecules, we would say this final state was more random, more mixed, than the initial state in which the red and white molecules were confined to specific regions. Another way to say this is in terms of ``disorder;'' there is more disorder in the final state than in the initial state. One view of entropy is thus that increases in entropy are connected with increases in randomness or disorder. This link can be made rigorous and is extremely useful in describing systems on a microscopic basis. While we do not have scope to examine this topic in depth, the purpose of this chapter is to make plausible the link between disorder and entropy through a statistical definition of entropy.

Next: 7.2 Microscopic and Macroscopic Up: 7. Entropy on the Previous: 7. Entropy on the Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.8 Some Overall Comments Up: 6. Applications of the Previous: 6.6 Entropy and Unavailable Contents Index

6.7 Examples of Lost Work in Engineering Processes


1. Lost work in Adiabatic Throttling: Entropy and Stagnation Pressure Changes

Figure 6.8: Adiabatic Throttling

2. A process we have encountered before is adiabatic throttling of a gas, by a valve or other device as shown in Figure 6.8. The velocity is denoted by . There is no shaft work and no heat transfer and the flow is steady. Under these conditions we can use the first law for a control volume (the Steady Flow Energy Equation) to make a statement about the conditions upstream and downstream of the valve:

3. 4. where is the stagnation enthalpy, corresponding to a (possibly fictitious) state with zero , even if

velocity. The stagnation enthalpy is the same at stations 1 and 2 if the flow processes are not reversible.

5. For a perfect gas with constant specific heats, and relation between the static and stagnation temperatures is:

. The

6.

7. where is the speed of sound and is the Mach number, . In deriving this result, use has only been made of the first law, the equation of state, the speed of sound, and the definition of the Mach number. Nothing has yet been specified about whether the process of stagnating the fluid is reversible or irreversible.

8. When we define the stagnation pressure, however, we do it with respect to isentropic deceleration to the zero velocity state. For an isentropic process

9.

10. The relation between static and stagnation pressures is

11.

Figure 6.9: Static and stagnation pressures and temperatures

12. The stagnation state is defined by

. In

addition, . The static and stagnation states are shown in - coordinates in Figure 6.9. 13. Stagnation pressure is a key variable in propulsion and power systems. To see why, we examine the relation between stagnation pressure, stagnation temperature, and entropy. The form of the combined first and second law that uses enthalpy is

(6..8)

14.

Figure 6.10:Stagnation and static states

15. This holds for small changes between any thermodynamic states and we can apply it to a situation in which we consider differences between stagnation states, say one state having properties and the other having

properties (see Figure 6.10). The corresponding static states are also indicated. Because the entropy is the same at static and stagnation conditions, needs no subscript. Writing (6.8) in terms of stagnation conditions yields

16. 17. Both sides of the above are perfect differentials and can be integrated as

18.

19. For a process with , the stagnation enthalpy, and hence the stagnation temperature, is constant. In this situation, the stagnation pressure is related directly to the entropy as,

(6..9)

20.

Figure 6.11:Losses reflected in changes in stagnation pressure when

21. Figure 6.11 shows this relation on a - diagram. We have seen that the entropy is related to the loss, or irreversibility. The stagnation pressure plays the role of an indicator of loss if the stagnation temperature is constant. The utility is that it is the stagnation pressure (and temperature) which are directly measured, not the entropy. The throttling process is a representation of flow through inlets, nozzles, stationary turbomachinery blades, and the use of stagnation pressure as a measure of loss is a practice that has widespread application. 22. Equation (6.9) can be put in several useful approximate forms. First, we note that for aerospace applications we are (hopefully!) concerned with low loss devices, so that the stagnation pressure change is small compared to the inlet level of stagnation pressure,

23. 24. Expanding the logarithm (using ),

26. or

25.

27. 28. Another useful form is obtained by dividing both sides by and taking the limiting forms ). Doing

of the expression for stagnation pressure in the limit of low Mach number ( this, we find:

29. 30. The quantity on the right can be interpreted as the change in the ``Bernoulli constant'' for incompressible (low Mach number) flow. The quantity on the left is a non-dimensional entropy change parameter, with the term now representing the loss of mechanical energy associated with the change in stagnation pressure.

31. To summarize:
1. For many applications the stagnation temperature is constant and the change in stagnation pressure is a direct measure of the entropy increase. 2. Stagnation pressure is the quantity that is actually measured so that linking it to entropy (which is not measured) is useful.

3. We can regard the throttling process as a ``free expansion'' at constant temperature from the initial stagnation pressure to the final stagnation pressure. We thus know that, for the process, the work we need to do to bring the gas back to the initial state is , which is the ``lost work'' per unit mass.

Muddy Points Why do we find stagnation enthalpy if the velocity never equals zero in the flow? (MP 6.13) Why does remain constant for throttling? (MP 6.14)

32. Adiabatic Efficiency of a Propulsion System Component (Turbine)

Figure 6.12: Schematic of turbine and associated thermodynamic representation in coordinates

33. A schematic of a turbine and the accompanying thermodynamic diagram are given in Figure 6.12. There is a pressure and temperature drop through the turbine and it produces work. There is no heat transfer so the expressions that describe the overall shaft work and the shaft work per unit mass are
(6..10)

(6..11)

34. 35. If the difference in the kinetic energy at inlet and outlet can be neglected, Equation (6.11) reduces to 36. 37. The adiabatic efficiency of the turbine is defined as

38. 39. The performance of the turbine can be represented in an - plane (similar to a plane for a perfect gas with constant specific heats) as shown in Figure 6.12. From the figure the adiabatic efficiency is

40.

41. The adiabatic efficiency can therefore be written as

42. 43. The non-dimensional term represents the departure from isentropic (reversible) processes and hence a loss. The quantity is the enthalpy difference for two states along a constant pressure line (see diagram). From the combined first and second laws, for a constant pressure process, small changes in enthalpy are related to the entropy change by , or approximately, 44. 45. The adiabatic efficiency can thus be approximated as

47. The quantity irreversibility.

46. represents a useful figure of merit for fluid machinery inefficiency due to

48.

Muddy Points 49. How do you tell the difference between shaft work/power and flow work in a turbine, both conceptually and mathematically? (MP 6.15)
50. Isothermal Expansion with Friction

Figure 6.13: Isothermal expansion with friction

51. In a more general look at the isothermal expansion, we now drop the restriction to frictionless processes. As seen in Figure 6.13, work is done to overcome friction. If the kinetic energy of the piston is negligible, a balance of forces tells us that
52.

53. During the expansion, the piston and the walls of the container will heat up because of the friction. The heat will be (eventually) transferred to the atmosphere; all frictional work ends up as heat transferred to the surrounding atmosphere. 54.

55. The amount of heat transferred to the atmosphere due to the frictional work only is thus,

56.

57. The entropy change of the atmosphere (considered as a heat reservoir) due to the

frictional work is 58. The engine operates in a cycle and the entropy change for the complete cycle is zero (because entropy is a state variable). Therefore,

59. 60. The total entropy change is

61.

62. Suppose we had an ideal reversible engine working between these same two temperatures, which extracted the same amount of heat, temperature reservoir, and rejected heat of magnitude reservoir. The work done by this reversible engine is
63. 64. For the reversible engine the total entropy change over a cycle is

, from the high to the low temperature

65.

66. Combining the expressions for work and for the entropy changes,
67. 68. The entropy change for the irreversible cycle can therefore be written as

69. 70. The difference in work that the two cycles produce is proportional to the entropy that is

generated during the cycle: 71. 72. The second law states that the total entropy generated is greater than zero for an irreversible process, so that the reversible work is greater than the actual work of the irreversible cycle.

73. An ``engine effectiveness,'' , can be defined as the ratio of the actual work obtained divided by the work that would have been delivered by a reversible engine operating between the two temperatures and :

74. 75. The departure from a reversible process is directly reflected in the entropy change and the decrease in engine effectiveness.

76.

Muddy Points 77. Why does ? (MP 6.17) 78. In discussing the terms ``closed system'' and ``isolated system,'' can you assume that you are discussing a cycle or not? (MP 6.18) 79. Does a cycle process have to have ? (MP 6.19) 80. In a real heat engine, with friction and losses, why is ? (MP 6.20)
81. Propulsive Power and Entropy Flux

still 0 if

The final example in this section combines a number of ideas presented in this subject and in Unified in the development of a relation between entropy generation and power needed to propel a vehicle. Figure 6.15 shows an aerodynamic shape (airfoil) moving through the atmosphere at a constant velocity. A coordinate system fixed to the vehicle has been adopted so that we see the airfoil as fixed and the air far away from the airfoil moving at a velocity . Streamlines of the flow have been sketched, as has the velocity distribution at station ``0'' far upstream and station ``d'' far downstream.

The airfoil has a wake, which mixes with the surrounding air and grows in the downstream direction. The extent of the wake is also indicated. Because of the lower velocity in the wake the area between the stream surfaces is larger downstream than upstream.

Figure 6.15: Airfoil with wake and control volume for analysis of propulsive power requirement

We use a control volume description and take the control surface to be defined by the two stream surfaces and two planes at station 0 and station d. This is useful in simplifying the analysis because there is no flow across the stream surfaces. The area of the downstream plane control surface is broken into , which is the area

outside the wake and , which is the area occupied by wake fluid, i.e., fluid that has suffered viscous losses. The control surface is also taken far enough away from the vehicle so that the static pressure can be considered uniform. For fluid which is not in the wake (no viscous forces), the momentum equation is

Uniform static pressure therefore implies uniform velocity, so that on

the velocity is

equal to the upstream value, . The downstream velocity profile is actually continuous, as indicated. It is approximated in the analysis as a step change to make the algebra a bit simpler. (The conclusions apply to the more general velocity profile as well and we would just need to use integrals over the wake instead of the algebraic expressions below.)

The equation expressing mass conservation for the control volume is


(6..12)

The vertical face of the control surface is far downstream of the body. By this station, the wake fluid has had much time to mix and the velocity in the wake is close to the free stream

value,

. We can thus write, (6..13)

(We chose our control surface so the condition

was upheld.)

The integral momentum equation (control volume form of the momentum equation) can be used to find the drag on the vehicle:
(6..14)

There is no pressure contribution in Eq. (6.14) because the static pressure on the control surface is uniform. Using the form given for the wake velocity and expanding the terms in the momentum equation we obtain
(6..15)

The last term in the right hand side of the momentum equation, , is small by virtue of the choice of control surface and we can neglect it. Doing this and grouping the terms on the right hand side of Eq. (6.15) in a different manner, we have

The terms in the square brackets on both sides of this equation are the continuity equation multiplied by . They thus sum to zero leaving the curly bracketed terms as (6..16)

The wake mass flow is . All this flow has a velocity defect (compared to the free stream) of , so that the defect in flux of momentum (the mass flow in the wake times the velocity defect) is, to first order in ,

The combined first and second law gives us a means of relating the entropy and velocity:

The pressure is uniform ( ) at the downstream station. There is no net shaft work or heat transfer to the wake so that the mass flux of stagnation enthalpy is constant. We can also approximate that the condition of constant stagnation enthalpy holds locally on all streamlines. Applying both of these to the combined first and second law yields

For the present situation,

, so that (6..17)

In Equation (6.17) the upstream temperature is used because differences between wake quantities and upstream quantities are small at the downstream control station. The entropy can be related to the drag as (6..18)

The quantity is the entropy flux (mass flux times the entropy increase per unit mass; in the general case we would express this by an integral over the locally varying wake velocity and density). The power needed to propel the vehicle is the product of , . From Eq. (6.18), this can be related to the entropy flux in the wake to yield a compact expression for the propulsive power needed in terms of

the wake entropy flux: (6..19)

This amount of work is dissipated per unit time in connection with sustaining the vehicle motion. Equation (6.19) is another demonstration of the relation between lost work and entropy generation, in this case manifested as power that needs to be supplied because of dissipation in the wake.

Muddy Points Is it safe to say that entropy is the tendency for a system to go into disorder? (MP 6.21)

Thermodynamics and Propulsion

Next: 5.2 Axiomatic Statements of Up: 5. The Second Law Previous: 5. The Second Law Contents Index

5.1 Concept and Statements of the Second Law (Why do we need a second law?)
The unrestrained expansion, or the temperature equilibration of the two bricks, are familiar processes. Suppose you are asked whether you have ever seen the reverse of these processes take place? Do two bricks at a medium temperature ever go to a state where one is hot and one is cold? Will the gas in the unrestrained expansion ever spontaneously return to occupying only the left side of the volume? Experience hints that the answer is no. However, both these processes, unfamiliar though they may be, are compatible with the first law. In other words the first law does not prohibit their occurrence. There thus

must be some other ``great principle'' that describes the direction of natural processes, that tells us which first law compatible processes will not be observed. This is contained in the second law. Like the first law, it is a generalization from an enormous amount of observation. There are several ways in which the second law of thermodynamics can be stated. Listed below are three that are often encountered. As described in class (and as derived in almost every thermodynamics textbook), although the three may not appear to have much connection with each other, they are equivalent.
1. No process is possible whose sole result is the absorption of heat from a reservoir and the conversion of this heat into work. [Kelvin-Planck statement of the second law]

Figure 5.1: This is not possible (Kelvin-Planck)

2. No process is possible whose sole result is the transfer of heat from a cooler to a hotter body. [Clausius statement of the second law]

Figure 5.2: For

, this is not possible (Clausius)

3. There exists for every system in equilibrium a property called entropy, , which is a thermodynamic property of a system. For a reversible process, changes in this property are given by

The entropy change of any system and its surroundings, considered together, is positive and approaches zero for any process which approaches reversibility.

For an isolated system, i.e., a system that has no interaction with the surroundings, changes in the system have no effect on the surroundings. In this case, we need to consider the system only, and the first and second laws become:

For an isolated system the total energy ( ) is constant. The entropy can only increase or, in the limit of a reversible process, remain constant. The limit, or , represents the best that can be done. In thermodynamics, propulsion, and power generation systems we often compare performance to this limit to measure how close to ideal a given process is. All of these statements are equivalent, but 3 gives a direct, quantitative measure of the departure from reversibility. Entropy is not a familiar concept and it may be helpful to provide some additional rationale for its appearance. If we look at the first law,

the term on the left is a function of state, while the two terms on the right are not. For a simple compressible substance, however, we can write the work done in a reversible process as , so that

Two out of the three terms in this equation are expressed in terms of state variables. It seems plausible that we ought to be able to express the third term using state variables as well, but what are the appropriate variables? If so, the term should perhaps be viewed as analogous to where the parentheses denote an intensive state variable and the square brackets denote an extensive state variable. The second law tells

us that the intensive variable is the temperature, , and the extensive state variable is the entropy, . The first law for a simple compressible substance in terms of state variables is thus
(5..1)

Because Eq. 5.1 includes the second law, it is referred to as the combined first and second law. Because it is written in terms of state variables, it is true for all processes, not just reversible ones. We summarize below some attributes of entropy:
1. Entropy is a function of the state of the system and can be found if any two properties of the system are known, e.g. or or . 2. is an extensive variable. The entropy per unit mass, or specific entropy, is . 3. The units of entropy are Joules per degree Kelvin (J/K). The units for specific entropy are J/K-kg. 4. For a system, , where the numerator is the heat given to the system and the denominator is the temperature of the system at the location where the heat is received. 5. for pure work transfer.

Muddy Points

Why is

always true? (MP 5.1)

What makes

different than

? (MP 5.2)

Next: 5.2 Axiomatic Statements of Up: 5. The Second Law Previous: 5. The Second Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.6 Entropy and Unavailable Up: 6. Applications of the Previous: 6.4 Brayton Cycle in Contents Index

6.5 Irreversibility, Entropy Changes, and ``Lost Work''


Consider a system in contact with a heat reservoir during a reversible process. If there is heat absorbed by the reservoir at temperature , the change in entropy of the

reservoir is . In general, reversible processes are accompanied by heat exchanges that occur at different temperatures. To analyze these, we can visualize a sequence of heat reservoirs at different temperatures so that during any infinitesimal portion of the cycle there will not be any heat transferred over a finite temperature difference.

During any infinitesimal portion, heat one of the reservoirs which is at change of the system is . If

will be transferred between the system and is absorbed by the system, the entropy

The entropy change of the reservoir is

The total entropy change of system plus surroundings is

This is also true if there is a quantity of heat rejected by the system.

The conclusion is that for a reversible process, no change occurs in the total entropy produced, i.e., the entropy of the system plus the entropy of the

surroundings:

Figure 6.7:Irreversible and reversible state changes

We now carry out the same type of analysis for an irreversible process, which takes the system between the same specified states as in the reversible process. This is shown schematically in Figure 6.7, with and denoting the irreversible and reversible processes. In the irreversible process, the system receives heat . The change in internal energy for the irreversible process is and does work

For the reversible process

Because the state change is the same in the two processes (we specified that it was), the change in internal energy is the same. Equating the changes in internal energy in the above two expressions yields

The subscript ``actual'' refers to the actual process (which is irreversible). The entropy change associated with the state change is

(6..3)

If the process is not reversible, we obtain less work (see IAW notes) than in a reversible process, , so that for the irreversible process,

There is no equality between the entropy change

and the quantity

for an

irreversible process. The equality is only applicable for a reversible process.

The change in entropy for any process that leads to a transformation between an initial state ``a'' and a final state ``b'' is therefore

where is the heat exchanged in the actual process. The equality only applies to a reversible process.

The difference

represents work we could have obtained, but did not. It . In terms of this quantity we can write,

is referred to as lost work and denoted by

(6..4)

The content of Equation (6.4) is that the entropy of a system can be altered in two ways: (i) through heat exchange and (ii) through irreversibilities. The lost work ( in Equation (6.4)) is always greater than zero, so the only way to decrease the entropy of a system is through heat transfer. To apply the second law we consider the total entropy change (system plus surroundings). If the surroundings are a reservoir at temperature , with which the system exchanges heat,

The total entropy change is

The quantity (

) is the entropy generated due to irreversibility.

Yet another way to state the distinction we are making is


(6..5 )

The lost work is also called dissipation and noted entropy change of the system becomes:

. Using this notation, the infinitesimal

or

Equation (6.5) can also be written as a rate equation, (6..6)

Either of Equation (6.5) or (6.6) can be interpreted to mean that the entropy of the system, affected by two factors: the flow of heat

, is

and the appearance of additional entropy, denoted by

, due to irreversibility6.1. This additional entropy is zero when the process is reversible and always positive when the process is irreversible. Thus, one can say that the system developssources which create entropy during an irreversible process. The second law asserts that sinks of entropy are impossible in nature, which is a more graphic way of saying that and are positive definite (always greater than zero), or zero in the special case of reversible processes.

The term

which is associated with heat transfer to the system, can be interpreted as a flux of entropy. The boundary is crossed by heat and the ratio of this heat flux to temperature can be defined as a flux of entropy. There are no restrictions on the sign of this quantity, and we can say that this flux either contributes towards, or drains away, the system's entropy. During a reversible process, only this flux can affect the entropy of the system. This terminology suggests that we interpret entropy as a kind of weightless fluid, whose quantity is conserved (like that of matter) during a reversible process. During an irreversible process, however, this fluid is not conserved; it cannot disappear, but rather is created by sources throughout the system. While this interpretation should not be taken too literally, it provides an easy mode of expression and is in the same category of concepts such as those associated with the phrases ``flux of energy'' or ``sources of heat.'' In fluid mechanics, for example, this graphic language is very effective and there should be no objections to copying it in thermodynamics.

Muddy Points Do we ever see an absolute variable for entropy? So far, we have worked with deltas only (MP 6.8)

I am confused as to

as opposed to

.(MP 6.9)

For irreversible processes, how can we calculate

if not equal to

?(MP 6.10)

Next: 6.6 Entropy and Unavailable Up: 6. Applications of the Previous: 6.4 Brayton Cycle in Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 7.3 A Statistical Definition Up: 7. Entropy on the Previous: 7.1 Entropy Change in Contents Index

7.2 Microscopic and Macroscopic Descriptions of a System


The microscopic description of a system is the complete description of each particle in this system. In the above example, the microscopic description of the gas would be the list of the state of each molecule: position and velocity in this problem. It would require a great deal of data for this description; there are roughly molecules in a cube of air one centimeter on a side at room temperature and pressure. The macroscopic description, which is in terms of a few (two!) properties is thus far more accessible and useable for engineering applications, although it is restricted to equilibrium states. To address the description of entropy on a microscopic level, we need to state some results concerning microscopic systems. These results and the computations and arguments below are taken almost entirely from the excellent discussion in Chapter 6 of Engineering Thermodynamics by Reynolds and Perkins7.1. For a given macroscopic system, there are many microscopic states. A key idea from quantum mechanics is that the states of atoms, molecules, and entire systems are discretely quantized. This means that a system of particles under certain constraints, like being in a box of a specified size, or having a fixed total energy, can exist in a finite number of allowed microscopic states. This number can be very big, but it is finite. The microstates of the system keep changing with time from one quantum state to another as molecules move and collide with one another. The probability for the system to be in a particular quantum state is defined by its quantum-state probability . The set of the is called the distribution of probability. The sum of the probabilities of all the allowed quantum states must be unity, hence for any time ,

(7..2)

When the system reaches equilibrium, the individual molecules still change from one quantum state to another. In equilibrium, however, the system state does not change with time; so the probabilities for the different quantum states are independent of time. This distribution is then called the equilibrium distribution, and the probability can be viewed as the fraction of time a system spends in the quantum state. In what follows, we limit consideration to equilibrium states. We can get back to macroscopic quantities from the microscopic description using the

probability distribution. For instance, the macroscopic energy of the system would be the weighted average of the successive energies of the system (the energies of the quantum states); the energies are weighted by the relative time the system spends in the corresponding microstates. In terms of probabilities, the average energy, , is

(7..3)

where

is the energy of a quantum state.

The probability distribution provides information on the randomness of the equilibrium quantum states. For example, suppose the system can only exist in three states (1, 2 and 3). If the distribution probability is

the system is in quantum state 1 and there is no randomness. If we were asked what quantum state the system is in, we would be able to say it is always in state 1. If the distribution were

or

the randomness would not be zero and would be equal in both cases. We would be more uncertain about the instantaneous quantum state than in the first situation.

Maximum randomness corresponds to the case where the three states are equally probable:

In this case, we can only guess the instantaneous state with

probability.

Next: 7.3 A Statistical Definition Up: 7. Entropy on the Previous: 7.1 Entropy Change in Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.6 Entropy and Unavailable Up: 6. Applications of the Previous: 6.4 Brayton Cycle in Contents Index

6.5 Irreversibility, Entropy Changes, and ``Lost Work''


Consider a system in contact with a heat reservoir during a reversible process. If there is heat absorbed by the reservoir at temperature , the change in entropy of the

reservoir is . In general, reversible processes are accompanied by heat exchanges that occur at different temperatures. To analyze these, we can visualize a sequence of heat reservoirs at different temperatures so that during any infinitesimal portion of the cycle there will not be any heat transferred over a finite temperature difference.

During any infinitesimal portion, heat one of the reservoirs which is at change of the system is . If

will be transferred between the system and is absorbed by the system, the entropy

The entropy change of the reservoir is

The total entropy change of system plus surroundings is

This is also true if there is a quantity of heat rejected by the system.

The conclusion is that for a reversible process, no change occurs in the total entropy produced, i.e., the entropy of the system plus the entropy of the surroundings: .

Figure 6.7:Irreversible and reversible state changes

We now carry out the same type of analysis for an irreversible process, which takes the system between the same specified states as in the reversible process. This is shown schematically in Figure 6.7, with and denoting the irreversible and reversible processes. In the irreversible process, the system receives heat . The change in internal energy for the irreversible process is and does work

For the reversible process

Because the state change is the same in the two processes (we specified that it was), the change in internal energy is the same. Equating the changes in internal energy in the above two expressions yields

The subscript ``actual'' refers to the actual process (which is irreversible). The entropy change associated with the state change is (6..3)

If the process is not reversible, we obtain less work (see IAW notes) than in a reversible process, , so that for the irreversible process,

There is no equality between the entropy change and the quantity for an irreversible process. The equality is only applicable for a reversible process.

The change in entropy for any process that leads to a transformation between an initial state ``a'' and a final state ``b'' is therefore

where is the heat exchanged in the actual process. The equality only applies to a reversible process.

The difference

represents work we could have obtained, but did not. It . In terms of this quantity we can write,

is referred to as lost work and denoted by

(6..4)

The content of Equation (6.4) is that the entropy of a system can be altered in two ways: (i) through heat exchange and (ii) through irreversibilities. The lost work ( in Equation (6.4)) is always greater than zero, so the only way to decrease the entropy of a system is through heat transfer. To apply the second law we consider the total entropy change (system plus surroundings). If the surroundings are a reservoir at temperature , with which the system exchanges heat,

The total entropy change is

The quantity (

) is the entropy generated due to irreversibility.

Yet another way to state the distinction we are making is


(6..5 )

The lost work is also called dissipation and noted entropy change of the system becomes:

. Using this notation, the infinitesimal

or

Equation (6.5) can also be written as a rate equation, (6..6)

Either of Equation (6.5) or (6.6) can be interpreted to mean that the entropy of the system,

, is

affected by two factors: the flow of heat

and the appearance of additional entropy, denoted by

, due to irreversibility6.1. This additional entropy is zero when the process is reversible and always positive when the process is irreversible. Thus, one can say that the system developssources which create entropy during an irreversible process. The second law asserts that sinks of entropy are impossible in nature, which is a more graphic way of saying that and are positive definite (always greater than zero), or zero in the special case of reversible processes.

The term

which is associated with heat transfer to the system, can be interpreted as a flux of entropy. The boundary is crossed by heat and the ratio of this heat flux to temperature can be defined as a flux of entropy. There are no restrictions on the sign of this quantity, and we can say that this flux either contributes towards, or drains away, the system's entropy. During a reversible process, only this flux can affect the entropy of the system. This terminology suggests that we interpret entropy as a kind of weightless fluid, whose quantity is conserved (like that of matter) during a reversible process. During an irreversible process, however, this fluid is not conserved; it cannot disappear, but rather is created by sources throughout the system. While this interpretation should not be taken too literally, it provides an easy mode of expression and is in the same category of concepts such as those associated with the phrases ``flux of energy'' or ``sources of heat.'' In fluid mechanics, for example, this graphic language is very effective and there should be no objections to copying it in thermodynamics.

Muddy Points Do we ever see an absolute variable for entropy? So far, we have worked with deltas only (MP 6.8)

I am confused as to

as opposed to

.(MP 6.9)

For irreversible processes, how can we calculate

if not equal to

?(MP 6.10)

Next: 6.6 Entropy and Unavailable Up: 6. Applications of the Previous: 6.4 Brayton Cycle in Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 5.6 Muddiest Points on Up: 5. The Second Law Previous: 5.4 Entropy Changes in Contents Index

5.5 Calculation of Entropy Change in Some Basic Processes


1. Heat transfer from, or to, a heat reservoir.

A heat reservoir (Figure 5.3) is a constant temperature heat source or sink. Because the temperature is uniform, there is no heat transfer across a finite temperature difference and the heat exchange is reversible. From the definition of entropy ( ),

where is the heat into the reservoir (defined here as positive if heat flows into the reservoir.)

Figure 5.3: Heat transfer from/to a heat reservoir

2. Heat transfer between two heat reservoirs

The entropy change of the two reservoirs in Figure 5.4 is the sum of the entropy

change of each. If the high temperature reservoir is at reservoir is at , the total entropy change is

and the low temperature

Figure 5.4: Heat transfer between two reservoirs

The second law says that the entropy change must be equal to or greater than zero. This corresponds to the statement that heat must flow from the higher temperature source to the lower temperature source. This is one of the statements of the second law given in Section 5.1.

Muddy Points In the single reservoir example, why can the entropy decrease? (MP 5.6) Why does the entropy of a heat reservoir change if the temperature stays the same? (MP 5.7) How can the heat transfer from or to a heat reservoir be reversible? (MP 5.8) How can (MP 5.9) be less than zero in any process? Doesn't entropy always increase?

If the same

for a reservoir, could you add ? (MP 5.10)

to any size reservoir and still get

3. Possibility of obtaining work from a single heat reservoir

We can regard the process proposed in Figure 5.5 as the absorption of heat, , by a device or system, operating in a cycle, rejecting no heat, and producing work. The total entropy change is the sum of the change in the reservoir, the system or device, and the surroundings. The entropy change of the reservoir is . The entropy change of the device is zero, because we are considering a complete cycle (return to initial state) and entropy is a function of state. The surroundings receive work only so the entropy change of the surroundings is zero. The total entropy change is

Figure 5.5: Work from a single heat reservoir

The total entropy change in the proposed process is thus less than zero,

which is not possible. The second law thus tells us that we cannot get work from a single reservoir only. The ``only'' is important; it means without any other changes occurring. This is the other statement of the second law we saw in Section 5.1.

Muddy Points What is the difference between the isothermal expansion of a piston and the (forbidden) production of work using a single reservoir? (MP 5.11) For the ``work from a single heat reservoir'' example, how do we know there is no ? (MP 5.12) ? I thought that the whole thing about cycles

How does a cycle produce zero

was an entropy that the designers try to minimize. (MP 5.13)


4. Entropy changes in the ``hot brick problem''

[Temperature equalization of two bricks]

[Reservoirs used in

reversible state transformations] Figure 5.6: The ``Hot Brick'' Problem

5. We can examine in a more quantitative manner the changes that occurred when we put the two bricks together, as depicted in Figure 5.6(a). The process by which the two bricks come to the same temperature is not a reversible one, so we need to devise a reversible path. To do this imagine a large number of heat reservoirs at varying temperatures spanning the range , as in Figure 5.6(b). The bricks are put in contact with them sequentially to raise the temperature of one and lower the temperature of the other in a reversible manner. The heat exchange at any of these steps is brick, the entropy change is: . For the high temperature

7. where cold brick,

6. is the heat capacity of the brick (J/kg). This quantity is less than zero. For the

8. 9. The entropy change of the two bricks is

10. 11. The process is not reversible. 12. Difference between the free expansion and the reversible isothermal expansion of an ideal gas

The essential difference between the free expansion in an insulated enclosure and the reversible isothermal expansion of an ideal gas can also be captured clearly in terms of entropy changes. For a state change from initial volume and temperature , to final volume and (the same) temperature , the

entropy change is

or, making use of the equation of state and the fact that process,

for an isothermal

This is the entropy change that occurs for the free expansion as well as for the isothermal reversible expansion processes - entropy changes are state changes and the two system final and end states are the same for both processes.

For the free expansion:

There is no change in the entropy of the surroundings because there is no interaction between the system and the surroundings. The total entropy change is therefore,

There are several points to note from this result:


1. 2. so the process is not reversible. ; the equality between and is only for a

reversible process. 3. There is a direct connection between the work needed to restore the system to the original state and the entropy change:

The quantity has a physical meaning as ``lost work'' in the sense of work which we lost the opportunity to utilize. We will make this connection stronger in Chapter 6.

For the reversible isothermal expansion: The entropy is a state variable so the entropy change of the system is the same as before. In this case, however, heat is transferred to the system from the surroundings ( ) so that

The heat transferred from the surroundings, however, is equal to the heat received by the system: .

The total change in entropy (system plus surroundings) is therefore

The reversible process has zero total change in entropy.

Muddy Points On the example of free expansion versus isothermal expansion, how do we know that the pressure and volume ratios are the same? We know for each that . (MP 5.14) and

Where did

come from? (MP 5.15)

Next: 5.6 Muddiest Points on Up: 5. The Second Law Previous: 5.4 Entropy Changes in Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 5.5 Calculation of Entropy Up: 5. The Second Law Previous: 5.3 Combined First and Contents Index

5.4 Entropy Changes in an Ideal Gas


[VW, S & B: 6.5- 6.6, 7.1] Many aerospace applications involve flow of gases (e.g., air) and we thus examine the entropy relations for ideal gas behavior. The starting point is form (a) of the combined first and second law,

For an ideal gas,

. Thus

Using the equation of state for an ideal gas ( change as an expression with only exact differentials:

), we can write the entropy

(5..2)

We can think of Equation (5.2) as relating the fractional change in temperature to the fractional change of volume, with scale factors and ; if the volume increases without a proportionate decrease in temperature (as in the case of an adiabatic free expansion), then increases. Integrating Equation (5.2) between two states ``1'' and ``2'':

For a perfect gas with constant specific heats

In non-dimensional form (using

(5..3)

Equation 5.3 is in terms of specific quantities. For

moles of gas,

This expression gives entropy change in terms of temperature and volume. We can develop an alternative form in terms of pressure and volume, which allows us to examine an assumption we have used. The ideal gas equation of state can be written as

Taking differentials of both sides yields

Using the above equation in Eq. (5.2), and making use of the relations ; , we find

or

Integrating between two states 1 and 2

(5..4)

Using both sides of (5.4) as exponents we obtain

(5..5)

Equation (5.5) describes a general process. For the specific situation in which , i.e., the entropy is constant, we recover the expression . It was stated that this expression applied to a reversible, adiabatic process. We now see, through use of the second law, a deeper meaning to the expression, and to the concept of a reversible adiabatic process, in that both are characteristics of a constant entropy, or isentropic, process.

Muddy Points Why do you rewrite the entropy change in terms of ? (MP 5.4)

What is the difference between isentropic and adiabatic? (MP 5.5)

Next: 5.5 Calculation of Entropy Up: 5. The Second Law Previous: 5.3 Combined First and Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's All Contents Index Subsections

1.2.1 The Continuum Model 1.2.2 The Concept of a ``System'' 1.2.3 The Concept of a ``State'' 1.2.4 The Concept of ``Equilibrium'' 1.2.5 The Concept of a ``Process'' 1.2.6 Quasi-Equilibrium Processes 1.2.7 Equations of state

1.2 Definitions and Fundamental Ideas of Thermodynamics


As with all sciences, thermodynamics is concerned with the mathematical modeling of the real world. In order that the mathematical deductions are consistent, we need some precise definitions of the basic concepts. The following is a discussion of some of the concepts we will need. Several of these will be further amplified in the lectures and in other handouts. If you need additional information or examples concerning these topics, they are described clearly and in-depth in (SB&VW). They are also covered, although in a less detailed manner, in Chapters 1 and 2 of the book by Van Ness.

1.2.1 The Continuum Model


Matter may be described at a molecular (or microscopic) level using the techniques of statistical mechanics and kinetic theory. For engineering purposes, however, we want ``averaged'' information, i.e., a macroscopic, not a microscopic, description. There are two reasons for this. First, a microscopic description of an engineering device may produce too much information to manage. For example, pressure contains of air at standard temperature and molecules (VW, S & B:2.2), each of which has a position and a

velocity. Typical engineering applications involve more than molecules. Second, and more importantly, microscopic positions and velocities are generally not useful for determining how macroscopic systems will act or react unless, for instance, their total effect is integrated. We therefore neglect the fact that real substances are composed of discrete molecules and model matter from the start as a smoothed-outcontinuum. The information we have about a continuum represents the microscopic information averaged over a volume. Classical thermodynamics is concerned only with continua.

1.2.2 The Concept of a ``System''


A thermodynamic system is a quantity of matter of fixed identity, around which we can

draw a boundary (see Figure 1.3 for an example). The boundaries may be fixed or moveable. Work or heat can be transferred across the system boundary. Everything outside the boundary is the surroundings. When working with devices such as engines it is often useful to define the system to be an identifiable volume with flow in and out. This is termed a control volume. An example is shown in Figure 1.5. A closed system is a special class of system with boundaries that matter cannot cross. Hence the principle of the conservation of mass is automatically satisfied whenever we employ a closed system analysis. This type of system is sometimes termed a control mass.

Figure 1.3: Piston (boundary) and gas (system)

Figure 1.4: Boundary around electric motor (system)

Figure 1.5: Sample control volume

1.2.3 The Concept of a ``State''


The thermodynamic state of a system is defined by specifying values of a set of

measurable properties sufficient to determine all other properties. For fluid systems, typical properties are pressure, volume and temperature. More complex systems may require the specification of more unusual properties. As an example, the state of an electric battery requires the specification of the amount of electric charge it contains. Properties may be extensive or intensive. Extensive properties are additive. Thus, if the system is divided into a number of sub-systems, the value of the property for the whole system is equal to the sum of the values for the parts. Volume is an extensive property. Intensive properties do not depend on the quantity of matter present. Temperature and pressure are intensive properties. Specific properties are extensive properties per unit mass and are denoted by lower case letters. For example:

Specific properties are intensive because they do not depend on the mass of the system.

The properties of a simple system are uniform throughout. In general, however, the properties of a system can vary from point to point. We can usually analyze a general system by sub-dividing it (either conceptually or in practice) into a number of simple systems in each of which the properties are assumed to be uniform. It is important to note that properties describe states only when the system is in equilibrium.

Muddy Points Specific properties (MP 1.1) What is the difference between extensive and intensive properties? (MP 1.2)

1.2.4 The Concept of ``Equilibrium''


The state of a system in which properties have definite, unchanged values as long as external conditions are unchanged is called an equilibrium state.

[Mechanical Equilibrium]

[Thermal

Equilibrium] Figure 1.6: Equilibrium

A system in thermodynamic equilibrium satisfies:


1. mechanical equilibrium (no unbalanced forces) 2. thermal equilibrium (no temperature differences) 3. chemical equilibrium.

1.2.5 The Concept of a ``Process''


If the state of a system changes, then it is undergoing a process. The succession of states through which the system passes defines the path of the process. If, at the end of the process, the properties have returned to their original values, the system has undergone a cyclic process or a cycle. Note that even if a system has returned to its original state and completed a cycle, the state of the surroundings may have changed.

1.2.6 Quasi-Equilibrium Processes


We are often interested in charting thermodynamic processes between states on thermodynamic coordinates. Recall from the end of Section 1.2.3, however, that properties define a state only when a system is in equilibrium. If a process involves finite, unbalanced forces, the system can pass through non-equilibrium states, which we cannot treat. An extremely useful idealization, however, is that only ``infinitesimal'' unbalanced forces exist, so that the process can be viewed as taking place in a series of ``quasi-equilibrium'' states. (The term quasi can be taken to mean ``as if;'' you will see it used in a number of contexts such as quasi-one-dimensional, quasi-steady, etc.) For this to be true the process must be slow in relation to the time needed for the system to come to equilibrium internally. For a gas at conditions of interest to us, a given molecule can undergo roughly molecular collisions per second, so that, if ten collisions are needed to come to equilibrium, the equilibration time is on the order of seconds. This is generally much shorter than the time scales associated with the bulk properties of the flow (say the time needed for a fluid particle to move some significant fraction of the length of the device of interest). Over a large range of parameters, therefore, it is a very good approximation to view the thermodynamic processes as consisting of such a succession of equilibrium states, which we can chart. [VW, S& B: 2.3-2.4] The figures below demonstrate the use of thermodynamics coordinates to plot isolines, lines along which a property is constant. They include constant temperature lines, or isotherms, on a - diagram, constant volume lines, or isochors on a -

diagram, and constant pressure lines, or isobars, on a

diagram for an ideal gas.

Real substances may have phase changes (water to water vapor, or water to ice, for example), which we can also plot on thermodynamic coordinates. We will see such phase changes plotted and used for liquid-vapor power generation cycles in Chapter 8. A preview is given in Figure 1.15 at the end of this chapter.

Figure: Figure:

diagram diagram

Figure: - diagram Figure 1.7: Thermodynamics coordinates and isolines for an ideal gas

1.2.7 Equations of state


It is an experimental fact that two properties are needed to define the state of any pure substance in equilibrium or undergoing a steady or quasi-steady process. [VW, S & B: 3.1, 3.3]. Thus for a simple compressible gas like air,

where

is the volume per unit mass,

. In words, if we know

and

we know

, etc.

Any of these is equivalent to an equation , which is known as an equation of state. The equation of state for an ideal gas, which is a very good approximation to real gases at conditions that are typically of interest for aerospace applications1.2, is

where

is the volume per mol of gas and .

is the ``Universal Gas

Constant,''

A form of this equation which is more useful in fluid flow problems is obtained if we divide by the molecular weight, :

where R is

, which has a different value for different gases due to the different molecular .

weights. For air at room conditions,

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's All Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's All Contents Index Subsections

1.2.1 The Continuum Model 1.2.2 The Concept of a ``System'' 1.2.3 The Concept of a ``State'' 1.2.4 The Concept of ``Equilibrium'' 1.2.5 The Concept of a ``Process'' 1.2.6 Quasi-Equilibrium Processes 1.2.7 Equations of state

1.2 Definitions and Fundamental Ideas of Thermodynamics


As with all sciences, thermodynamics is concerned with the mathematical modeling of the real world. In order that the mathematical deductions are consistent, we need some precise definitions of the basic concepts. The following is a discussion of some of the concepts we will need. Several of these will be further amplified in the lectures and in other handouts. If you need additional information or examples concerning these topics,

they are described clearly and in-depth in (SB&VW). They are also covered, although in a less detailed manner, in Chapters 1 and 2 of the book by Van Ness.

1.2.1 The Continuum Model


Matter may be described at a molecular (or microscopic) level using the techniques of statistical mechanics and kinetic theory. For engineering purposes, however, we want ``averaged'' information, i.e., a macroscopic, not a microscopic, description. There are two reasons for this. First, a microscopic description of an engineering device may produce too much information to manage. For example, pressure contains of air at standard temperature and molecules (VW, S & B:2.2), each of which has a position and a

velocity. Typical engineering applications involve more than molecules. Second, and more importantly, microscopic positions and velocities are generally not useful for determining how macroscopic systems will act or react unless, for instance, their total effect is integrated. We therefore neglect the fact that real substances are composed of discrete molecules and model matter from the start as a smoothed-outcontinuum. The information we have about a continuum represents the microscopic information averaged over a volume. Classical thermodynamics is concerned only with continua.

1.2.2 The Concept of a ``System''


A thermodynamic system is a quantity of matter of fixed identity, around which we can draw a boundary (see Figure 1.3 for an example). The boundaries may be fixed or moveable. Work or heat can be transferred across the system boundary. Everything outside the boundary is the surroundings. When working with devices such as engines it is often useful to define the system to be an identifiable volume with flow in and out. This is termed a control volume. An example is shown in Figure 1.5. A closed system is a special class of system with boundaries that matter cannot cross. Hence the principle of the conservation of mass is automatically satisfied whenever we employ a closed system analysis. This type of system is sometimes termed a control mass.

Figure 1.3: Piston (boundary) and gas (system)

Figure 1.4: Boundary around electric motor (system)

Figure 1.5: Sample control volume

1.2.3 The Concept of a ``State''


The thermodynamic state of a system is defined by specifying values of a set of measurable properties sufficient to determine all other properties. For fluid systems, typical properties are pressure, volume and temperature. More complex systems may require the specification of more unusual properties. As an example, the state of an electric battery requires the specification of the amount of electric charge it contains. Properties may be extensive or intensive. Extensive properties are additive. Thus, if the system is divided into a number of sub-systems, the value of the property for the whole system is equal to the sum of the values for the parts. Volume is an extensive property. Intensive properties do not depend on the quantity of matter present. Temperature and pressure are intensive properties. Specific properties are extensive properties per unit mass and are denoted by lower case letters. For example:

Specific properties are intensive because they do not depend on the mass of the system.

The properties of a simple system are uniform throughout. In general, however, the properties of a system can vary from point to point. We can usually analyze a general system by sub-dividing it (either conceptually or in practice) into a number of simple systems in each of which the properties are assumed to be uniform.

It is important to note that properties describe states only when the system is in equilibrium.

Muddy Points Specific properties (MP 1.1) What is the difference between extensive and intensive properties? (MP 1.2)

1.2.4 The Concept of ``Equilibrium''


The state of a system in which properties have definite, unchanged values as long as external conditions are unchanged is called an equilibrium state.

[Mechanical Equilibrium]

[Thermal

Equilibrium] Figure 1.6: Equilibrium

A system in thermodynamic equilibrium satisfies:


1. mechanical equilibrium (no unbalanced forces) 2. thermal equilibrium (no temperature differences) 3. chemical equilibrium.

1.2.5 The Concept of a ``Process''


If the state of a system changes, then it is undergoing a process. The succession of states through which the system passes defines the path of the process. If, at the end of the process, the properties have returned to their original values, the system has undergone a cyclic process or a cycle. Note that even if a system has returned to its original state and completed a cycle, the state of the surroundings may have changed.

1.2.6 Quasi-Equilibrium Processes


We are often interested in charting thermodynamic processes between states on thermodynamic coordinates. Recall from the end of Section 1.2.3, however, that properties define a state only when a system is in equilibrium. If a process involves finite, unbalanced forces, the system can pass through non-equilibrium states, which we cannot treat. An extremely useful idealization, however, is that only ``infinitesimal'' unbalanced forces exist, so that the process can be viewed as taking place in a series of ``quasi-equilibrium'' states. (The term quasi can be taken to mean ``as if;'' you will see it used in a number of contexts such as quasi-one-dimensional, quasi-steady, etc.) For this to be true the process must be slow in relation to the time needed for the system to come to equilibrium internally. For a gas at conditions of interest to us, a given molecule can undergo roughly molecular collisions per second, so that, if ten collisions are needed to come to equilibrium, the equilibration time is on the order of seconds. This is generally much shorter than the time scales associated with the bulk properties of the flow (say the time needed for a fluid particle to move some significant fraction of the length of the device of interest). Over a large range of parameters, therefore, it is a very good approximation to view the thermodynamic processes as consisting of such a succession of equilibrium states, which we can chart. [VW, S& B: 2.3-2.4] The figures below demonstrate the use of thermodynamics coordinates to plot isolines, lines along which a property is constant. They include constant temperature lines, or isotherms, on a - diagram, constant volume lines, or isochors on a diagram, and constant pressure lines, or isobars, on a - diagram for an ideal gas. Real substances may have phase changes (water to water vapor, or water to ice, for example), which we can also plot on thermodynamic coordinates. We will see such phase changes plotted and used for liquid-vapor power generation cycles in Chapter 8. A preview is given in Figure 1.15 at the end of this chapter.

Figure: Figure:

diagram diagram

Figure: - diagram Figure 1.7: Thermodynamics coordinates and isolines for an ideal gas

1.2.7 Equations of state


It is an experimental fact that two properties are needed to define the state of any pure substance in equilibrium or undergoing a steady or quasi-steady process. [VW, S & B: 3.1, 3.3]. Thus for a simple compressible gas like air,

where

is the volume per unit mass,

. In words, if we know

and

we know

, etc.

Any of these is equivalent to an equation , which is known as an equation of state. The equation of state for an ideal gas, which is a very good approximation to real gases at conditions that are typically of interest for aerospace applications1.2, is

where

is the volume per mol of gas and .

is the ``Universal Gas

Constant,''

A form of this equation which is more useful in fluid flow problems is obtained if we divide by the molecular weight, :

where R is

, which has a different value for different gases due to the different molecular .

weights. For air at room conditions,

Next: 1.3 Changing the State Up: 1. Introduction to Thermodynamics Previous: 1.1 What it's All Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 12.4 Multistage Axial Compressors Up: 12. Energy Exchange with Previous: 12.2 Conservation of Angular Contents Index

12.3 The Euler Turbine Equation


The Euler turbine equation relates the power added to or removed from the flow, to characteristics of a rotating blade row. The equation is based on the concepts of conservation of angular momentum and conservation of energy. We will work with the model of the blade row shown in Figure 12.2.

Figure 12.2: Control volume for Euler Turbine Equation.

Applying conservation of angular momentum, we note that the torque, , must be equal to the time rate of change of angular momentum in a streamtube that flows through the device

This is true whether the blade row is rotating or not. The sign matters (i.e. angular momentum is a vector - positive means it is spinning in one direction, negative means it is spinning in the other direction). So depending on how things are defined, there can be positive and negative torques, and positive and negative angular momentum. In Figure 9.2, torque is positive when -- the same sense as the angular velocity.

If the blade row is moving, then work is done on/by the fluid. The work per unit time, or power, , is the torque multiplied by the angular velocity, :

If torque and angular velocity are of like sign, work is being done on the fluid (a compressor). If torque and angular velocity are of opposite sign work is being extracted from the fluid (a turbine). Here is another approach to the same idea:

If the tangential velocity increases across a blade row (where positive tangential velocity is defined in the same direction as the rotor motion) then work is added to the flow (this happens in a compressor). If the tangential velocity decreases across a blade row (where positive tangential velocity is defined in the same direction as the rotor motion) then work is removed from the flow (this happens in a turbine).

From the steady flow energy equation,

with

Then equating this expression of conservation of energy with our expression from conservation of angular momentum, we arrive at:

or for a perfect gas with

, (12..3)

Equation (12.3) is called the Euler Turbine Equation. It relates the temperature ratio (and hence the pressure ratio) across a turbine or compressor to the rotational speed and the change in momentum per unit mass. Note that the velocities used in this equation are what we will later call absolute frame velocities (as opposed to relative frame velocities).

If angular momentum increases across a blade row, then on the fluid (a compressor). If angular momentum decreases across a blade row, then by the fluid (a turbine).

and work was done

and work was done

Next: 12.4 Multistage Axial Compressors Up: 12. Energy Exchange with Previous: 12.2 Conservation of Angular Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements of Contents Index

5.3 Combined First and Second Law Expressions


The first law, written in a form that is always true:

For reversible processes only, work or heat may be rewritten as

Substitution leads to other forms of the first law true for reversible processes only:

(If the substance has other work modes, e.g., stress, strain,

where

is a pressure-like quantity, and

is a volume-like quantity.)

Substituting for both

and

in terms of state variables,

The above is always true because it is a relation between properties and is now independent of process.

In terms of specific quantities:

The combined first and second law expressions are often more usefully written in terms of the enthalpy, or specific enthalpy, ,

Or, since

In terms of enthalpy (rather than specific enthalpy) the relation is

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements of Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

2.1 First Law of Thermodynamics


[VW, S & B: 2.6]

Observation leads to the following two assertions:


1. There exists for every system a property called energy, . The system energy can be considered as a sum of internal energy, kinetic energy, potential energy, and chemical energy. 1. Like the Zeroth Law, which defined a useful property, ``temperature,'' the First Law defines a useful property called ``energy.'' 2. The two new terms (compared to what you have seen in physics and dynamics, for example) are the internal energy and the chemical energy. For most situations in this class, we will neglect the chemical energy. We will generally not, however, neglect the internal energy, . It arises from the random or disorganized motion of molecules in the system, as shown in Figure 2.1. Since this molecular motion is primarily a function of temperature, the internal energy is sometimes called ``thermal energy.''

Figure 2.1: Random motion is the physical basis for internal energy

3. The internal energy,

, is a function of the state of the system.

Thus , or , or . Recall that for pure substances the entire state of the system is specified if any two properties are specified. (We will discuss the equations that relate the internal energy to these other variables as the course progresses.)
2. The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system, (2..1)

3. 4. where is the energy of the system,

is the heat input to the system, and is the work done by the system. (thermal energy) + 1. Like the Zeroth Law, the First Law describes the behavior of the new property [VW, S& B: Chapter 5]. 2. The equation can also be written on a per unit mass basis

3. 4. In many situations the potential energy, kinetic energy, and chemical energy of the system are constant or not important. Then

5. 6. and

7.

8. Note that

and

are not functions of state, but

, which arises from

molecular motion (see above), depends only on the state of the system; does not depend on how the system got to that state. We therefore have the striking result that:

9.

10. Sometimes this difference is emphasized by writing the First Law in differential form,
(2..2)

11. 12. where the symbol `` '' is used to denote that these are not exact differentials but rather are dependent on path. 13. Note that the signs are important:

is defined to be positive if it is transferred to the system; thus the numerical value we substitute for will be positive if heat is transferred to the system from the surroundings, and negative if heat is transferred from the system to the surroundings. [VW, S & B: 4.7-4.8] is defined to be positive if it is done by the system (see Section 1.3); thus the numerical value we substitute for will be positive if the system is doing work, and negative if work is being done on the system. [VW, S& B: 4.1-4.4] ,

14. For quasi-static processes we can substitute

15.

To give an example of where the first law is applied, consider the device shown in Figure 2.2. We heat a gas, it expands against a weight, some force (pressure times area) is applied over a distance, and work is done. The change in energy of the system supplies the connection between the heat added and work done. We will spend most of the course dealing with various applications of the first law -- in one form or another.

Figure 2.2: The change in energy of a system relates the heat added to the work done

The form of the first law we have given here is sometimes called the ``control mass'' form, because it is well suited to dealing with systems of a fixed mass. We will see in Section 2.5 that this form can be written for a control volume with mass flow in and mass flow out (like a jet engine for example). We will call this the ``control volume'' form of the first law [VW, S & B: 5.8-5.12].

Muddy Points What are the conventions for work and heat in the first law? (MP 2.1) When does ? (MP 2.2)

Next: 2.2 Corollaries of the Up: 2. The First Law Previous: 2. The First Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 2.6 Muddiest Points on Up: 2. The First Law Previous: 2.4 Specific Heats Contents Index Subsections

2.5.1 Conservation of mass 2.5.2 Conservation of energy 2.5.3 Stagnation Temperature and Stagnation Enthalpy o 2.5.3.1 Frame dependence of stagnation quantities o 2.5.3.2 Example o 2.5.3.3 Steady Flow Energy Equation in terms of Stagnation Enthalpy

2.5.4 Example Applications of the First Law of Thermodynamics o 2.5.4.1 Tank Filling o 2.5.4.2 Flow through a rocket nozzle o 2.5.4.3 Power to drive a gas turbine compressor

2.5 Control volume form of the conservation laws


[VWB&S, 6.1, 6.2]

The thermodynamic laws (as well as Newton's laws) are for a system, a specific quantity of matter. More often, in propulsion and power problems, we are interested in what happens in a fixed volume, for example a rocket motor or a jet engine through which mass is flowing at a certain rate. We may also be interested in the rates of heat and work into and out of a system. For this reason, the control volume form of the system laws is of great importance. A schematic of the difference is shown in Figure 2.8. Rather than focus on a particle of mass which moves through the engine, it is more convenient to focus on the volume occupied by the engine. This requires us to use the control volume form of the thermodynamic laws, developed below.

Figure 2.8: Control volume and system for flow through a propulsion device

2.5.1 Conservation of mass


For the control volume shown, the rate of change of mass inside the volume is given by the difference between the mass flow rate in and the mass flow rate out. For a single flow coming in and a single flow coming out this is

If the mass inside the control volume changes with time it is because some mass is added or some is taken out. In the special case of a steady flow, , therefore

Figure 2.9: A control volume used to track mass flows

2.5.2 Conservation of energy


The first law of thermodynamics can be written as a rate equation:

where

To derive the first law as a rate equation for a control volume we proceed as with the mass conservation equation. The physical idea is that any rate of change of energy in the control volume must be caused by the rates of energy flow into or out of the volume. The heat transfer and the work are already included and the only other contribution must be associated with the mass flow in and out, which carries energy with it. Figure 2.10 shows two schematics of this idea. The desired form of the equation will be

[Simple]

[More

General] Figure 2.10: Schematic diagrams illustrating terms in the energy equation for a simple and a more general control volume

The fluid that enters or leaves has an amount of energy per unit mass given by

where is the fluid velocity relative to some coordinate system, and we have neglected chemical energy. In addition, whenever fluid enters or leaves a control volume there is a work term associated with the entry or exit. We saw this in Section 2.3, example 1, and the present derivation is essentially an application of the ideas presented there. Flow exiting at station ``e'' must push back the surrounding fluid, doing work on it. Flow entering the volume at station ``i'' is pushed on by, and receives work from the surrounding air. The rate of flow work at exit is given by the product of the pressure times the exit area times the rate at which the external flow is ``pushed back.'' The latter, however, is equal to the volume per unit mass times the rate of mass flow. Put another way, in a time , the work done on the surroundings by the flow at the exit station is

The net rate of flow work is

Including all possible energy flows (heat, shaft work, shear work, piston work, etc.), the first law can then be written as:

where includes the sign associated with the energy flow. If heat is added or work is done on the system then the sign is positive, if work or heat are extracted from the system then the sign is negative. NOTE: this is consistent with , where is the work done by the system on the environment, thus work is flowing out of the system.

We can then combine the specific internal energy term, work term, , to make the enthalpy appear:

, in

and the specific flow

Thus, the first law can be written as:

For most of the applications in this course, there will be no shear work and no piston work. Hence, the first law for a control volume will be most often used as:

(2..10)

Note how our use of enthalpy has simplified the rate of work term. In writing the control volume form of the equation we have assumed only one entering and one leaving stream, but this could

be generalized to any number of inlet and exit streams.

In the special case of a steady-state flow,

Applying this to Equation 2.10 produces a form of the ``Steady Flow Energy Equation'' (SFEE),

(2..11)

which has units of Joules per second. We could also divide by the mass flow to produce

which has units of Joules per second per kilogram. For problems of interest in aerospace applications the velocities are high and the term that is associated with changes in the elevation is small. From now on, we will neglect the terms unless explicitly stated.

Muddy Points What is shaft work? (MP 2.5) What distinguishes shaft work from other works? (MP 2.6) Definition of a control volume (MP 2.7)

2.5.3 Stagnation Temperature and Stagnation Enthalpy


Suppose that our steady flow control volume is a set of streamlines describing the flow up to the nose of a blunt object, as in Figure 2.11.

Figure 2.11: Streamlines and a stagnation region; a control volume can be drawn between the dashed streamlines and points 1 and 2

The streamlines are stationary in space, so there is no external work done on the fluid as it flows. If there is also no heat transferred to the flow (adiabatic), then the steady flow energy equation becomes

The quantity that is conserved is defined as the stagnation temperature,

or

where is the Mach number2.5. The stagnation temperature is the temperature that the fluid would reach if it were brought to zero speed by a steady adiabatic process with no external work. Note that for any steady, adiabatic flow with no external work, the stagnation temperature is constant.

It is also convenient to define the stagnation enthalpy,

which allows us to write the Steady Flow Energy Equation in a simpler form as

Note that for a quasi-static adiabatic process

so we can write

and define the relationship between stagnation pressure and static pressure as

where, the stagnation pressure is the pressure that the fluid would reach if it were brought to zero speed, via a steady, adiabatic, quasi-static process with no external work.

2.5.3.1 Frame dependence of stagnation quantities


An area of common confusion is the frame dependence of stagnation quantities. The stagnation temperature and stagnation pressure are the conditions the fluid would reach if it were brought to zero speed relative to some reference frame, via a steady adiabatic process with no external work (for stagnation temperature) or a steady, adiabatic, reversible process with no external work (for stagnation pressure). Depending on the

speed of the reference frame the stagnation quantities will take on different values. For example, consider a high speed reentry vehicle traveling through the still atmosphere, which is at temperature, . Let's place our reference frame on the vehicle and stagnate a fluid particle on the nose of the vehicle (carrying it along with the vehicle and thus essentially giving it kinetic energy). The stagnation temperature of the air in the vehicle frame is

where is the vehicle speed. The temperature the skin reaches (to first approximation) is the stagnation temperature and depends on the speed of the vehicle. Since re-entry vehicles travel fast, the skin temperature is much hotter than the atmospheric temperature. The atmospheric temperature, , is not frame dependent, but the stagnation temperature, , is.

The confusion comes about because is usually referred to as the static temperature. In common language this has a similar meaning as ``stagnation,'' but in fluid mechanics and thermodynamics static is used to label the thermodynamic properties of the gas ( etc.), and these are not frame dependent. , ,

Thus in our re-entry vehicle example, looking at the still atmosphere from the vehicle frame we see a stagnation temperature hotter than the atmospheric (static) temperature. If we look at the same still atmosphere from a stationary frame, the stagnation temperature is the same as the static temperature.

2.5.3.2 Example
For the case shown below, a jet engine is sitting motionless on the ground prior to take-off. Air is entrained into the engine by the compressor. The inlet can be assumed to be frictionless and adiabatic.

Figure 2.12: A stationary gas turbine drawing air in from the atmosphere

Considering the state of the gas within the inlet, prior to passage into the compressor, as

state (1), and working in the reference frame of the motionless airplane:

1. Is

greater than, less than, or equal to

The stagnation temperature of the atmosphere, , is equal to since it is moving the same speed as the reference frame (the motionless airplane). The steady flow energy equation tells us that if there is no heat or shaft work (the case for our adiabatic inlet) the stagnation enthalpy (and thus stagnation temperature for constant ) remains unchanged. Thus

2. Is

greater than, less than, or equal to

If then since the flow is moving at station 1 and therefore some of the total energy is composed of kinetic energy (at the expense of internal energy, thus lowering
3. Is

)
?

greater than, less than, or equal to

Equal, by the same argument as 1.


4. Is greater than, less than, or equal to ?

Less than, by the same argument as 2.

2.5.3.3 Steady Flow Energy Equation in terms of Stagnation Enthalpy


The form of the ``Steady Flow Energy Equation'' (SFEE) that we will most commonly use is Equation 2.11 written in terms of stagnation quantities, and neglecting chemical and potential energies,

The steady flow energy equation finds much use in the analysis of power and propulsion devices and other fluid machinery. Note the prominent role of enthalpy.

Muddy Points What is the difference between enthalpy and stagnation enthalpy? (MP 2.8)

2.5.4 Example Applications of the First Law of Thermodynamics


[VW, S& B: 6.4]

2.5.4.1 Tank Filling


Using what we have just learned we can attack the tank filling problem solved in Section 2.3.3 from an alternate point of view using the control volume form of the first law. In this problem the shaft work is zero, and the heat transfer, kinetic energy changes, and potential energy changes are neglected. In addition there is no exit mass flow.

Figure 2.13: A control volume approach to the tank filling problem

The control volume form of the first law is therefore

The equation of mass conservation is

Combining we have

Integrating from the initial time to the final time (the incoming enthalpy is constant) and using gives the result as before.

2.5.4.2 Flow through a rocket nozzle


A liquid bi-propellant rocket consists of a thrust chamber and nozzle and some means for

forcing the liquid propellants into the chamber where they react, converting chemical energy to thermal energy.

Figure 2.14: Flow through a rocket nozzle

Once the rocket is operating we can assume that all of the flow processes are steady, so it is appropriate to use the steady flow energy equation. Also, for now we will assume that the gas behaves as a perfect gas with constant specific heats, though in general this is a poor approximation. There is no external work, and we assume that the flow is adiabatic. We define our control volume as going between location , in the chamber, and location , at the exit, and then write the First Law as

or

Therefore

If we assume quasi-static, adiabatic expansion then

so

and , the conditions in the combustion chamber, are set by propellants, and external static pressure.

is the

2.5.4.3 Power to drive a gas turbine compressor


Consider for example the PW4084 pictured in Figure 2.15. The engine is designed to produce about 84,000 lbs of thrust at takeoff. The engine is a two-spool design. The fan and low pressure compressor are driven by the low pressure turbine. The high pressure compressor is driven by the high pressure turbine. We wish to find the total shaft work required to drive the compression system.

Figure 2.15: The Pratt and Whitney 4084 (drawing courtesy of Pratt and Whitney)

We define our control volume to encompass the compression system, from the front of the fan to the back of the fan and high pressure compressor, with the shaft cutting through the back side of the control volume. Heat transfer from the gas streams is negligible, so we write the First Law (steady flow energy equation) as:

For this problem we must consider two streams, the fan stream,

, and the core stream,

We obtain the temperature change by assuming that the compression process is quasi-static and adiabatic,

then

Substituting these values into the expression for the first law above, along with estimates of we obtain

Note that . If a car engine has , then the power needed to drive compressor is equivalent to 1,110 automobile engines. All of this power is generated by the low pressure and high pressure turbines.

Next: 2.6 Muddiest Points on Up: 2. The First Law Previous: 2.4 Specific Heats Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 15.4 First Law Analysis Up: 15. Generating Heat: Thermochemistry Previous: 15.2 Fuel-Air Ratio Contents Index

15.3 Enthalpy of formation

The systems we have worked with until now have been of fixed chemical composition. Because of this, we could use thermodynamic properties relative to an arbitrary base, since all comparisons could be made with respect to the chosen base. For example, the specific energy for steam. If there are no changes in composition, and only changes in properties of given substances, this is adequate. If there are changes in composition, however, we need to have a reference state so there is consistency for different substances. The convention used is that the reference state is a temperature of ( ) and a pressure of . (These are roughly room conditions.) At these reference conditions, the enthalpy of the elements (oxygen, hydrogen, nitrogen, carbon, etc.) is taken as zero. The results of a combustion process can be diagrammed as in Figure 15.1. The reactants enter at standard conditions; the combustion (reaction) takes place in the volume indicated. Downstream of the reaction zone there is an appropriate amount of heat transfer with the surroundings so that the products leave at the standard conditions. For the reaction of carbon and oxygen to produce is , the heat that has to be extracted

; this is heat that comes out of the control volume.

Figure 15.1: Constant pressure combustion

There is no shaft work done in the control volume and the first law for the control volume (SFEE) reduces to:

We can write this statement in the form (15..1)

In Eq. (15.1) the subscripts ``R'' and ``P'' on the summations refer to the reactants ( ) and products ( ) respectively. The subscripts on the mass flow rates and enthalpies refer to all of the components at inlet and at exit.

The relation in terms of mass flows can be written in molar form, which is often more convenient for reacting flow problems, by using the molecular weight, molar mass flow rate, component as , and molar enthalpy, , to define the

, for any individual ith (or eth)

The SFEE is, in these terms, (15..2)

The statements that have been made do not necessarily need to be viewed in the context of flow processes. Suppose we have one unit of and one unit of at the initial conditions and we . If so,

carry out a constant pressure reaction at ambient pressure,

since

. Combining terms,

or

In terms of the numbers of moles and the specific enthalpy this is (15..3)

The enthalpy of

, at

and

, with reference to a base where the enthalpy of . Values of the heat of

the elements is zero, is called the enthalpy of formation and denoted by formation for a number of substances are given in Table A.9 in SB&VW.

The enthalpies of the reactants and products for the formation of

are:

The enthalpy of

in any other state

is given by

These descriptions can be applied to any compound. For elements or compounds that exist in more than one state at the reference conditions (for example, carbon exists as diamond and as graphite), we also need to specify the state.

Note that there is a minus sign for the heat of formation. The heat transfer is out of the control volume and is thus negative by our convention. This means that .

Muddy Points Is the enthalpy of formation equal to the heat transfer out of the combustion during the formation reaction? (MP 15.4) Are the enthalpies of and (monoatomic hydrogen) both zero at ?

(MP 15.5)

Next: 15.4 First Law Analysis Up: 15. Generating Heat: Thermochemistry Previous: 15.2 Fuel-Air Ratio Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.8 Some Overall Comments Up: 6. Applications of the Previous: 6.6 Entropy and Unavailable Contents Index

6.7 Examples of Lost Work in Engineering Processes


1. Lost work in Adiabatic Throttling: Entropy and Stagnation Pressure Changes

Figure 6.8: Adiabatic Throttling

2. A process we have encountered before is adiabatic throttling of a gas, by a valve or other device as shown in Figure 6.8. The velocity is denoted by . There is no shaft work and no heat transfer and the flow is steady. Under these conditions we can use the first law for a control volume (the Steady Flow Energy Equation) to make a statement about the conditions upstream and downstream of the valve:

3. 4. where is the stagnation enthalpy, corresponding to a (possibly fictitious) state with zero , even if

velocity. The stagnation enthalpy is the same at stations 1 and 2 if

the flow processes are not reversible.

5. For a perfect gas with constant specific heats, and relation between the static and stagnation temperatures is:

. The

6.

7. where is the speed of sound and is the Mach number, . In deriving this result, use has only been made of the first law, the equation of state, the speed of sound, and the definition of the Mach number. Nothing has yet been specified about whether the process of stagnating the fluid is reversible or irreversible.

8. When we define the stagnation pressure, however, we do it with respect to isentropic deceleration to the zero velocity state. For an isentropic process

9.

10. The relation between static and stagnation pressures is

11.

Figure 6.9: Static and stagnation pressures and temperatures

12. The stagnation state is defined by addition, in - coordinates in Figure 6.9.

. In . The static and stagnation states are shown

13. Stagnation pressure is a key variable in propulsion and power systems. To see why, we examine the relation between stagnation pressure, stagnation temperature, and entropy. The form of the combined first and second law that uses enthalpy is

(6..8)

14.

Figure 6.10:Stagnation and static states

15. This holds for small changes between any thermodynamic states and we can apply it to a situation in which we consider differences between stagnation states, say one state having properties and the other having

properties (see Figure 6.10). The corresponding static states are also indicated. Because the entropy is the same at static and stagnation conditions, needs no subscript. Writing (6.8) in terms of stagnation conditions yields

16. 17. Both sides of the above are perfect differentials and can be integrated as

18.

19. For a process with , the stagnation enthalpy, and hence the stagnation temperature, is constant. In this situation, the stagnation pressure is related directly to the entropy as,

(6..9)

20.

Figure 6.11:Losses reflected in changes in stagnation pressure when

21. Figure 6.11 shows this relation on a - diagram. We have seen that the entropy is related to the loss, or irreversibility. The stagnation pressure plays the role of an indicator of loss if the stagnation temperature is constant. The utility is that it is the stagnation pressure (and temperature) which are directly measured, not the entropy. The throttling process is a representation of flow through inlets, nozzles, stationary turbomachinery blades, and the use of stagnation pressure as a measure of loss is a practice that has widespread application. 22. Equation (6.9) can be put in several useful approximate forms. First, we note that for aerospace applications we are (hopefully!) concerned with low loss devices, so that the stagnation pressure change is small compared to the inlet level of stagnation pressure,

23. 24. Expanding the logarithm (using ),

26. or

25.

27. 28. Another useful form is obtained by dividing both sides by and taking the limiting

forms of the expression for stagnation pressure in the limit of low Mach number ( ). Doing this, we find:

29.

30. The quantity on the right can be interpreted as the change in the ``Bernoulli constant'' for incompressible (low Mach number) flow. The quantity on the left is a non-dimensional entropy change parameter, with the term now representing the loss of mechanical energy associated with the change in stagnation pressure.

31. To summarize:
1. For many applications the stagnation temperature is constant and the change in stagnation pressure is a direct measure of the entropy increase. 2. Stagnation pressure is the quantity that is actually measured so that linking it to entropy (which is not measured) is useful. 3. We can regard the throttling process as a ``free expansion'' at constant temperature from the initial stagnation pressure to the final stagnation pressure. We thus know that, for the process, the work we need to do to bring the gas back to the initial state is , which is the ``lost work'' per unit mass.

Muddy Points Why do we find stagnation enthalpy if the velocity never equals zero in the flow? (MP 6.13) Why does remain constant for throttling? (MP 6.14)

32. Adiabatic Efficiency of a Propulsion System Component (Turbine)

Figure 6.12: Schematic of turbine and associated thermodynamic representation in coordinates

33. A schematic of a turbine and the accompanying thermodynamic diagram are given in Figure 6.12. There is a pressure and temperature drop through the turbine and it produces work. There is no heat transfer so the expressions that describe the overall shaft work and the shaft work per unit mass are
(6..10)

(6..11)

34. 35. If the difference in the kinetic energy at inlet and outlet can be neglected, Equation (6.11) reduces to 36. 37. The adiabatic efficiency of the turbine is defined as

38. 39. The performance of the turbine can be represented in an - plane (similar to a plane for a perfect gas with constant specific heats) as shown in Figure 6.12. From the figure the adiabatic efficiency is

40.

41. The adiabatic efficiency can therefore be written as

42. 43. The non-dimensional term represents the departure from isentropic (reversible) processes and hence a loss. The quantity is the enthalpy difference for two states along a constant pressure line (see diagram). From the combined first and second laws, for a constant pressure process, small changes in enthalpy are related to the entropy change by , or approximately, 44. 45. The adiabatic efficiency can thus be approximated as

47. The quantity to irreversibility.

46. represents a useful figure of merit for fluid machinery inefficiency due

48.

Muddy Points 49. How do you tell the difference between shaft work/power and flow work in a turbine, both conceptually and mathematically? (MP 6.15)
50. Isothermal Expansion with Friction

Figure 6.13: Isothermal expansion with friction

51. In a more general look at the isothermal expansion, we now drop the restriction to frictionless processes. As seen in Figure 6.13, work is done to overcome friction. If the kinetic energy of the piston is negligible, a balance of forces tells us that
52. 53. During the expansion, the piston and the walls of the container will heat up because of the friction. The heat will be (eventually) transferred to the atmosphere; all frictional work ends up as heat transferred to the surrounding atmosphere. 54.

55. The amount of heat transferred to the atmosphere due to the frictional work only is thus,

56.

57. The entropy change of the atmosphere (considered as a heat reservoir) due to the

frictional work is 58. The engine operates in a cycle and the entropy change for the complete cycle is zero (because entropy is a state variable). Therefore,

59. 60. The total entropy change is

61.

62. Suppose we had an ideal reversible engine working between these same two temperatures, which extracted the same amount of heat, , from the high

temperature reservoir, and rejected heat of magnitude to the low temperature reservoir. The work done by this reversible engine is

63. 64. For the reversible engine the total entropy change over a cycle is

65.

66. Combining the expressions for work and for the entropy changes,
67. 68. The entropy change for the irreversible cycle can therefore be written as

69. 70. The difference in work that the two cycles produce is proportional to the entropy that is generated during the cycle: 71. 72. The second law states that the total entropy generated is greater than zero for an irreversible process, so that the reversible work is greater than the actual work of the irreversible cycle.

73. An ``engine effectiveness,'' , can be defined as the ratio of the actual work obtained divided by the work that would have been delivered by a reversible engine operating between the two temperatures and :

74. 75. The departure from a reversible process is directly reflected in the entropy change and the decrease in engine effectiveness.

76.

Muddy Points 77. Why does ? (MP 6.17) 78. In discussing the terms ``closed system'' and ``isolated system,'' can you assume

that you are discussing a cycle or not? (MP 6.18) 79. Does a cycle process have to have ? (MP 6.19) 80. In a real heat engine, with friction and losses, why is ? (MP 6.20)
81. Propulsive Power and Entropy Flux

still 0 if

The final example in this section combines a number of ideas presented in this subject and in Unified in the development of a relation between entropy generation and power needed to propel a vehicle. Figure 6.15 shows an aerodynamic shape (airfoil) moving through the atmosphere at a constant velocity. A coordinate system fixed to the vehicle has been adopted so that we see the airfoil as fixed and the air far away from the airfoil moving at a velocity . Streamlines of the flow have been sketched, as has the velocity distribution at station ``0'' far upstream and station ``d'' far downstream. The airfoil has a wake, which mixes with the surrounding air and grows in the downstream direction. The extent of the wake is also indicated. Because of the lower velocity in the wake the area between the stream surfaces is larger downstream than upstream.

Figure 6.15: Airfoil with wake and control volume for analysis of propulsive power requirement

We use a control volume description and take the control surface to be defined by the two stream surfaces and two planes at station 0 and station d. This is useful in simplifying the analysis because there is no flow across the stream surfaces. The area of the downstream plane control surface is broken into , which is the area

outside the wake and , which is the area occupied by wake fluid, i.e., fluid that has suffered viscous losses. The control surface is also taken far enough away from the vehicle so that the static pressure can be considered uniform. For fluid which is not in the wake (no viscous forces), the momentum equation is

Uniform static pressure therefore implies uniform velocity, so that on

the velocity is

equal to the upstream value, . The downstream velocity profile is actually continuous, as indicated. It is approximated in the analysis as a step change to make the algebra a bit simpler. (The conclusions apply to the more general velocity profile as well and we would just need to use integrals over the wake instead of the algebraic expressions below.)

The equation expressing mass conservation for the control volume is


(6..12)

The vertical face of the control surface is far downstream of the body. By this station, the wake fluid has had much time to mix and the velocity in the wake is close to the free stream value, . We can thus write, (6..13)

(We chose our control surface so the condition

was upheld.)

The integral momentum equation (control volume form of the momentum equation) can be used to find the drag on the vehicle:
(6..14)

There is no pressure contribution in Eq. (6.14) because the static pressure on the control surface is uniform. Using the form given for the wake velocity and expanding the terms in the momentum equation we obtain
(6..15)

The last term in the right hand side of the momentum equation, , is small by virtue of the choice of control surface and we can neglect it. Doing this and grouping the terms on the right hand side of Eq. (6.15) in a different manner, we have

The terms in the square brackets on both sides of this equation are the continuity equation multiplied by . They thus sum to zero leaving the curly bracketed terms as (6..16)

The wake mass flow is . All this flow has a velocity defect (compared to the free stream) of , so that the defect in flux of momentum (the mass flow in the wake times the velocity defect) is, to first order in ,

The combined first and second law gives us a means of relating the entropy and velocity:

The pressure is uniform ( ) at the downstream station. There is no net shaft work or heat transfer to the wake so that the mass flux of stagnation enthalpy is constant. We can also approximate that the condition of constant stagnation enthalpy holds locally on all streamlines. Applying both of these to the combined first and second law yields

For the present situation,

, so that (6..17)

In Equation (6.17) the upstream temperature is used because differences between wake quantities and upstream quantities are small at the downstream control station. The entropy can be related to the drag as (6..18)

The quantity is the entropy flux (mass flux times the entropy increase per unit mass; in the general case we would express this by an integral over the locally varying wake velocity and density). The power needed to propel the vehicle is the product of , . From Eq. (6.18), this can be related to the entropy flux in the wake to yield a compact expression for the propulsive power needed in terms of the wake entropy flux: (6..19)

This amount of work is dissipated per unit time in connection with sustaining the vehicle motion. Equation (6.19) is another demonstration of the relation between lost work and entropy generation, in this case manifested as power that needs to be supplied because of dissipation in the wake.

Muddy Points Is it safe to say that entropy is the tendency for a system to go into disorder? (MP 6.21)

Thermodynamics and Propulsion

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index Subsections

11.6.1 Notation and station numbering 11.6.2 Ideal Assumptions 11.6.3 Ideal Ramjet o 11.6.3.1 Thrust o 11.6.3.2 Fuel Air Ratio
o o o

11.6.3.3 Specific impulse, 11.6.3.4 Representative performance values 11.6.3.5 Recapitulation

11.6.4 Turbojet Engine 11.6.5 Effect of Departures from Ideal Behavior o 11.6.5.1 Parameters reflecting design choices o 11.6.5.2 Parameters reflecting the ability to design and execute efficient components

11.6 Performance of Jet Engines


In Chapter 3 we represented a gas turbine engine using a Brayton cycle and derived expressions for efficiency and work as functions of the temperature at various points in the cycle. In this section we will perform further ideal cycle analysis to express the thrust and fuel efficiency of engines in terms of useful design variables, including design limits, flight conditions, and design choices. The expressions we develop will allow us to define a particular mission and then determine the optimum component characteristics (e.g. compressor, combustor, turbine) for an engine for a given mission. Note that ideal cycle analysis addresses only the thermodynamics of the airflow within the engine. It does not describe the details of the components (the blading, the rotational speed, etc.), but only the results the various components produce (e.g. pressure ratios, temperature ratios). In Chapter 12 we will look in greater detail at how some of the components (the turbine and the compressor)

produce these effects.

11.6.1 Notation and station numbering


Notation:

where the capital subscript

will refer to the turbine. Stagnation properties,

and

, are

more easily measured quantities than static properties ( and ). Thus, it is standard convention to express the performance of various components in terms of stagnation pressure and temperature ratios:

total or stagnation pressure ratio across component ( , total or stagnation temperature ratio across component (

, , ,

, ,

, ,

, ,

) ),

where , , , , and the engine we will use the numbers in Figure 11.12.

, . To identify individual stations within

Figure 11.12: Gas turbine engine station numbering.

11.6.2 Ideal Assumptions


1. Inlet/Diffuser: , (adiabatic, isentropic)

2. Compressor or fan: 3. Combustor/burner or afterburner:

, ,

4. Turbine: 5. Nozzle: , .

11.6.3 Ideal Ramjet


To get started with a simple example (no turbomachinery), we will reexamine the ideal ramjet, picking up where we left off in Section 3.7.3. (Note that we will continue to use station 5 as the exit station, consistent with Figure 3.25.)

11.6.3.1 Thrust
The coordinate system and control volume are chosen to be fixed to the ramjet. The thrust, , is given by:

where and are the exit and inlet flow velocities, respectively. The thrust can be put in terms of nondimensional parameters as follows:

Using

in the expression for stagnation pressure,

The ratios of stagnation pressure to static pressure at inlet and exit of the ramjet are

The ratios of stagnation to static pressure at exit and at inlet are the same, with the consequence that the inlet and exit Mach numbers are also the same.

To find the thrust we need to find the ratio of the temperature at exit and the temperature at inlet. This is given by

where

is the stagnation temperature ratio across the combustor (burner). The thrust is thus

11.6.3.2 Fuel Air Ratio


To find the Isp will will need to find the ramjet fuel-air ratio, the burner, as shown in Figure 11.13, we get: . Using a control volume around

From the steady flow energy equation:

The exit mass flow is not greatly different from the inlet mass flow, , because the fuel-air ratio is much less than unity (generally several percent). We thus neglect the difference between the mass flows and obtain

with

Thus, the fuel-air ratio is

The fuel-air ratio,

, depends on the fuel properties (

), the desired flight parameters (

), the ramjet performance (

), and the temperature of the atmosphere (

).

Figure 11.13: Control volume over the burner

11.6.3.3 Specific impulse,


The specific impulse for the ramjet is given by

The specific impulse can be written in terms of fuel properties and flight and vehicle characteristics as

We wish to explore the parameter dependency of the above expression, which is a complicated formula. How can we do this? What are the important effects of the different parameters? How do we best capture the ramjet performance behavior? To make effective comparisons, we need to add some additional information concerning the operational behavior. An important case to examine is when we have stoichiometric conditions and all the fuel burns (denoted by ). What happens in this situation as

the flight Mach number, , increases? is fixed so increases, but the maximum temperature does not increase much because of dissociation: the reaction does not go to completion at high temperature. A useful approximation is therefore to take km, constant for stoichiometric operation. In the stratosphere, from 10 to 30 . The maximum temperature ratio is

For the stoichiometric ramjet,

Using the expression for

, the specific impulse is

11.6.3.4 Representative performance values

Figure 11.14: Thrust per unit mass flow and specific impulse for ideal ramjet with stoichiometric combustion [Kerrebrock]

A plot of the performance of the stoichiometric ramjet is shown in Figure 11.14. The figure shows that for the parameters used, the best operating range of a hydrocarbonfueled ramjet is the stratosphere, . The parameters used are for hydrocarbons, such that , . in

11.6.3.5 Recapitulation
Between Section 3.7.3 and this section we have:

Examined the Brayton Cycle for ramjet propulsion,

Found Found Examined

as a function of

, and . , and

and the relation between and as a function of

Muddy Points What exactly is the specific impulse, How is found for rockets in space where rather than

, a measure of? (MP 11.1) ? (MP 11.2)

Why does industry use (MP 11.3)

? Is there an advantage to this?

Why isn't mechanical efficiency an issue with ramjets? (MP 11.4) How is thrust created in a ramjet? (MP 11.5)

What is the relation between value of ? (MP 11.6)

and the existence of the maximum

Why didn't we have a 2s point for the Brayton cycle with non-ideal components? (MP 11.7)

What is the variable

? (MP 11.8)

11.6.4 Turbojet Engine

Figure 11.15: Schematic with appropriate component notations added.

We now examine an engine with turbomachinery, as shown in Figure 11.15. Methodology:

1. Find thrust by finding in terms of , temperature ratios, etc. 2. Use a power balance to relate turbine parameters to compressor parameters 3. Use an energy balance across the combustor to relate the combustor temperature rise to the fuel flow rate and fuel energy content. First write-out the expressions for thrust and Isp:

where

is the fuel/air mass flow ratio,

and

Now we have to do a little algebra to manipulate these expressions into more useful forms. First we write an expression for the exit velocity:

Noting that

we can write

Thus (11..1)

which expresses the exit temperature as a function of the inlet temperature, the Mach number, and the temperature changes across each component. We now write the pressure at the exit in a similar manner:

Since

and

we write

and then equate this to our expression for the temperature in (11.1):

and

We now continue on the path to our expression for

Therefore

Now we have two steps left. First we write in terms of , by noting that they are related by the condition that the power used by the compressor is equal to the power extracted by the turbine. Second, we put the burner temperature ratio in terms of the exit temperature of the burner, ( or more specifically ) since this is the hottest point in the engine and is a frequent benchmark used for judging various designs.

The steady flow energy equation states

Assuming that the compressor and turbine are adiabatic, then

Since the turbine shaft is connected to the compressor shaft

Assuming the mass flow and specific heats are the same between compressor and turbine, this can be rewritten as

where

so

or

That was the first step relating the temperature rise across the turbine to that across the compressor. The remaining step is to write the temperature rise across the combustor in terms of .

and for an engine with an afterburner

Now substituting our expressions for , and into our expression for into the first expression we wrote for thrust, we get:

, and finally

This is what we were seeking, an expression for thrust in terms of important design parameters and flight parameters:

By adding and substracting the quantity

we may write this write in another form which is often used,

A recap of the important variables:

Our final step involves writing the specific impulse and other measures of efficiency in terms of these same parameters. We begin by writing the First Law across the combustor to relate the fuel flow rate and heating value of the fuel to the total enthalpy rise.

and

where again,

is the fuel/air mass flow ratio. The specific impulse becomes

where is expressed in terms of typical design parameters, flight conditions, and physical constants

Similarly, we can write our overall efficiency as

or

Using the definition from before, the ideal thermal efficiency is

and the propulsive efficiency can be found as

We can now use these equations to better understand the performance of a simple turbojet engine. We will use the following parameters (with ):

Table 11.1: Turbojet Mission Parameters Mach number Altitude Ambient Temp. Speed of sound 0 0.85 2.0 Sea level 12 km 18 km 288 K 217 K 217 K 340 m/s 295 m/s 295 m/s

Note it is more typical to work with the compressor pressure ratio ( temperature ratio (

) rather than the

) so we will substitute the isentropic relationship:

into the equations before plotting the results in Figures 11.16, 11.17, and 11.18.

Figure 11.16: Performance of an ideal turbojet engine as a function of flight Mach number.

Figure 11.17: Performance of an ideal turbojet engine as a function of compressor pressure ratio and flight Mach number.

Figure 11.18: Performance of an ideal turbojet engine as a function of compressor pressure ratio and turbine inlet temperature.

11.6.5 Effect of Departures from Ideal Behavior -- Real Cycle behavior


To conclude this chapter, we will now improve our estimates of cycle performance by including the effects of irreversibility. We will use the Brayton cycle as an example. What are the sources of non-ideal performance and departures from reversibility?

Losses (entropy production) in the compressor and the turbine. Stagnation pressure decrease in the combustor. Heat transfer.

We take into account here only irreversibility in the compressor and in the turbine.

Because of these irreversibilities, we need more work, (the changes in kinetic energy from inlet to exit of the compressor are neglected), to drive the compressor than in the ideal situation. We also get less work, , back from the turbine. The consequence, as can be inferred from Figure 11.19, is that the net work from the engine is less than in the cycle with ideal components.

Figure 11.19: Gas turbine engine (Brayton) cycle showing effect of departure from ideal behavior in compressor and turbine

To develop a quantitative description of the effect of these departures from reversible behavior, consider a perfect gas with constant specific heats and neglect kinetic energy at the inlet and exit of the turbine and compressor. We define the turbine adiabatic efficiency as

where is specified to be at the same pressure ratio as for the compressor, the compressor adiabatic efficiency:

. There is a similar metric

again for the same pressure ratio. Note that for the turbine the ratio is the actual work delivered divided by the ideal work, whereas for the compressor the ratio is the ideal work needed divided by the actual work required. These are not thermal efficiencies, but rather measures of the degree to which the compression and expansion approach the ideal processes.

We now wish to find the net work done in the cycle and the efficiency. The net work is given either by the difference between the heat received and rejected or the work of the compressor and turbine, where the convention is that heat received is positive and heat rejected is negative and work done is positive and work absorbed is negative.

The thermal efficiency is

We need to calculate

From the definition of

With

Similarly, by the definition

we can find

The thermal efficiency can now be found:

With

and

or

There are several non-dimensional parameters that appear in this expression for thermal efficiency. We list these in the two sections below and show their effects in accompanying figures.

11.6.5.1 Parameters reflecting design choices

11.6.5.2 Parameters reflecting the ability to design and execute efficient components

In addition to efficiency, net rate of work is a quantity we need to examine,

Putting this in a non-dimensional form,

[Non-dimensional work as a function of cycle pressure ratio for different values of turbine entry

temperature divided by compressor entry temperature] [Overall cycle efficiency as a function of pressure ratio for different values of turbine entry temperature

divided by compressor entry temperature] [Overall cycle efficiency as a function of cycle pressure ratio for different component

efficiencies] Figure 11.20: Non-dimensional power and efficiency for a non-ideal gas turbine engine [from Cumpsty, Jet Propulsion]

Trends in net power and efficiency are shown in Figure 11.20 for parameters typical of advanced civil engines. Some points to note in the figure:

For any , the optimum pressure ratio for maximum is not the highest that can be achieved, as it is for the ideal Brayton cycle. The ideal analysis is too idealized in this regard. The highest efficiency also occurs closer to the pressure ratio for maximum power than in the case of an ideal cycle. Choosing this as a design criterion will therefore not lead to the efficiency penalty inferred from ideal cycle analysis. There is a strong sensitivity to the component efficiencies. For example, for , the cycle efficiency is roughly two-thirds of the ideal value.

The maximum power occurs at a value of max

or pressure ratio

less than that for

(this trend is captured by ideal analysis). are strongly dependent on the maximum

The maximum power and maximum temperature, .

A scale diagram of a Brayton cycle with non-ideal compressor and turbine behaviors, in terms of temperature-entropy ( - ) and pressure-volume ( - ) coordinates is given below as Figure 11.21.

[]

[]

Figure 11.21: Scale diagram of non-ideal gas turbine cycle. Nomenclature is shown in the figure. Pressure ratio , , , compressor and turbine efficiencies [from Cumpsty, Jet Propulsion]

Muddy Points Isn't it possible for the mixing of two gases to go from the final state to the initial state? If you have two gases in a box, they should eventually separate by density, right? (MP 11.9)

How can

be the maximum turbine inlet temperature? (MP 11.10)

When there are losses in the turbine that shift the expansion in - diagram to the right, does this mean there is more work than ideal since the area is greater? (MP 11.11) For an afterburning engine, why must the nozzle throat area increase if the temperature of the fluid is increased? (MP 11.12)

Why doesn't the pressure in the afterburner go up if heat is added? (MP 11.13) Why is the flow in the nozzle choked? (MP 11.14) What's the point of having a throat if it creates a retarding force? (MP 11.15) Why isn't the stagnation temperature conserved in this steady flow? (MP 11.16)

Next: 11.7 Performance of Propellers Up: 11. Aircraft Engine Performance Previous: 11.5 Trends in thermal Contents Index

UnifiedTP UnifiedTP UnifiedTP

Thermodynamics and Propulsion

Next: 15.3 Enthalpy of formation Up: 15. Generating Heat: Thermochemistry Previous: 15.1 Fuels Contents Index

15.2 Fuel-Air Ratio


The reaction for aeroengine fuel at stoichiometric conditions is

On a molar basis, the ratio of fuel to air is

To find the ratio on a mass flow basis, which is the way in which the aeroengine industry discusses it, we need to ``weight'' the molar proportions by the molecular weight of the components. The fuel molecular weight is , the oxygen molecular weight . The fuel/air

is and the nitrogen molecular weight is (approximately) ratio on a mass flow basis is thus

If we used the actual constituents of air we would get 0.0667, a value about 0.5% different.

Muddy Points Do we always assume 100% complete combustion? How good an approximation is this? (MP 15.3)

Next: 15.3 Enthalpy of formation Up: 15. Generating Heat: Thermochemistry Previous: 15.1 Fuels Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 5.2 Axiomatic Statements of Up: 5. The Second Law Previous: 5. The Second Law Contents Index

5.1 Concept and Statements of the Second Law (Why do we need a second law?)
The unrestrained expansion, or the temperature equilibration of the two bricks, are familiar processes. Suppose you are asked whether you have ever seen the reverse of these processes take place? Do two bricks at a medium temperature ever go to a state where one is hot and one is cold? Will the gas in the unrestrained expansion ever spontaneously return to occupying only the left side of the volume? Experience hints that the answer is no. However, both these processes, unfamiliar though they may be, are compatible with the first law. In other words the first law does not prohibit their occurrence. There thus must be some other ``great principle'' that describes the direction of natural processes, that tells us which first law compatible processes will not be observed. This is contained in the second law. Like the first law, it is a generalization from an enormous amount of observation.

There are several ways in which the second law of thermodynamics can be stated. Listed below are three that are often encountered. As described in class (and as derived in almost every thermodynamics textbook), although the three may not appear to have much connection with each other, they are equivalent.
1. No process is possible whose sole result is the absorption of heat from a reservoir and the conversion of this heat into work. [Kelvin-Planck statement of the second law]

Figure 5.1: This is not possible (Kelvin-Planck)

2. No process is possible whose sole result is the transfer of heat from a cooler to a hotter body. [Clausius statement of the second law]

Figure 5.2: For

, this is not possible (Clausius)

3. There exists for every system in equilibrium a property called entropy, , which is a thermodynamic property of a system. For a reversible process, changes in this property are given by

The entropy change of any system and its surroundings, considered together, is positive and approaches zero for any process which approaches reversibility.

For an isolated system, i.e., a system that has no interaction with the surroundings, changes in the system have no effect on the surroundings. In this case, we need to consider the system only, and the first and second laws become:

For an isolated system the total energy ( ) is constant. The entropy can only increase or, in the limit of a reversible process, remain constant. The limit, or , represents the best that can be done. In thermodynamics, propulsion, and power generation systems we often compare performance to this limit to measure how close to ideal a given process is. All of these statements are equivalent, but 3 gives a direct, quantitative measure of the departure from reversibility. Entropy is not a familiar concept and it may be helpful to provide some additional rationale for its appearance. If we look at the first law,

the term on the left is a function of state, while the two terms on the right are not. For a simple compressible substance, however, we can write the work done in a reversible process as , so that

Two out of the three terms in this equation are expressed in terms of state variables. It seems plausible that we ought to be able to express the third term using state variables as well, but what are the appropriate variables? If so, the term should perhaps be viewed as analogous to where the parentheses denote an intensive state variable and the square brackets denote an extensive state variable. The second law tells us that the intensive variable is the temperature, , and the extensive state variable is the entropy, . The first law for a simple compressible substance in terms of state variables is thus
(5..1)

Because Eq. 5.1 includes the second law, it is referred to as the combined first and second law. Because it is written in terms of state variables, it is true for all processes, not just reversible ones. We summarize below some attributes of entropy:
1. Entropy is a function of the state of the system and can be found if any two properties of the system are known, e.g. or or . 2. is an extensive variable. The entropy per unit mass, or specific entropy, is . 3. The units of entropy are Joules per degree Kelvin (J/K). The units for specific entropy are J/K-kg. 4. For a system, , where the numerator is the heat given to the system and the denominator is the temperature of the system at the location where the heat is received. 5. for pure work transfer.

Muddy Points

Why is

always true? (MP 5.1)

What makes

different than

? (MP 5.2)

Next: 5.2 Axiomatic Statements of Up: 5. The Second Law Previous: 5. The Second Law Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements

of Contents Index

5.3 Combined First and Second Law Expressions


The first law, written in a form that is always true:

For reversible processes only, work or heat may be rewritten as

Substitution leads to other forms of the first law true for reversible processes only:

(If the substance has other work modes, e.g., stress, strain,

where

is a pressure-like quantity, and

is a volume-like quantity.)

Substituting for both

and

in terms of state variables,

The above is always true because it is a relation between properties and is now independent of process.

In terms of specific quantities:

The combined first and second law expressions are often more usefully written in terms of the enthalpy, or specific enthalpy, ,

Or, since

In terms of enthalpy (rather than specific enthalpy) the relation is

Next: 5.4 Entropy Changes in Up: 5. The Second Law Previous: 5.2 Axiomatic Statements of Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 19.3 Radiation Heat Transfer Up: 19. Radiation Heat Transfer Previous: 19.1 Ideal Radiators Contents Index

19.2 Kirchhoff's Law and ``Real Bodies''


Real bodies radiate less effectively than black bodies. The measurement of this is the emittance, , defined by

where

is radiation from the real body at

, and

is radiation from a black body at

Values of emittance vary greatly for different materials. They are near unity for rough surfaces such as ceramics or oxidized metals, and roughly 0.02 for polished metals or silvered reflectors. A table of emittances for different substances is given at the end of this section as Table 19.1, taken from the book by Lienhard. The level of the emittance can be related to the absorptance using the following arguments. Suppose we have a small non-black body in the cavity. The power absorbed per unit area is equal to . The power emitted is equal to . An energy balance gives . Thus

(19..1)

Equation (19.1), the relation

, is known as Kirchhoff's Law. It implies that good radiators are

good absorbers. It was derived for the case when and is not strictly true for all circumstances when the temperature of the body and the cavity are different, but it is true if , , so the absorptance and emittance are not functions of , are properties of the surface, . This situation .

describes a ``gray body.'' Also, since

Next: 19.3 Radiation Heat Transfer Up: 19. Radiation Heat Transfer Previous: 19.1 Ideal Radiators Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 6.5 Irreversibility, Entropy Changes, Up: 6. Applications of the Previous: 6.3 Representation of Thermodynamic Contents Index Subsections

6.4.1 Net work per unit mass flow in a Brayton cycle

6.4 Brayton Cycle in

Coordinates

The Brayton cycle has two reversible adiabatic (i.e., isentropic) legs and two reversible, constant pressure heat exchange legs. The former are vertical, but we need to define the shape of the latter. For an ideal gas, changes in specific enthalpy are related to changes in temperature by , so the shape of the cycle in an plane is the same as in

a - plane, with a scale factor of between the two. This suggests that a place to start is with the combined first and second law, which relates changes in enthalpy, entropy, and pressure:

On constant pressure curves

and

. The quantity desired is the derivative of . From the , this is

temperature, , with respect to entropy, , at constant pressure: combined first and second law, and the relation between and

(6..2)

The derivative is the slope of the constant pressure legs of the Brayton cycle on a plane. For a given ideal gas (specific ) the slope is positive and increases as .

We can also plot the Brayton cycle in an - plane. This has advantages because changes in enthalpy directly show the work of the compressor and turbine and the heat added and rejected. The slope of the constant pressure legs in the - plane is .

Note that the similarity in the shapes of the cycles in - and - planes is true for ideal gases only. As we will see when we examine two-phase cycles, the shapes look quite different in these two planes when the medium is not an ideal gas.

Figure 6.4: Ideal Brayton cycle as composed of many elementary Carnot cycles [Kerrebrock]

Plotting the cycle in - coordinates also allows another way to address the evaluation of the Brayton cycle efficiency which gives insight into the relations between Carnot cycle efficiency and efficiency of other cycles. As shown in Figure 6.4, we can break up the Brayton cycle into many small Carnot cycles. The `` '' Carnot cycle has an efficiency of

where the indicated lower temperature is the heat rejection temperature for that elementary cycle and the higher temperature is the heat absorption temperature for that cycle. The upper and lower curves of the Brayton cycle, however, have constant pressure. All of the elementary Carnot cycles therefore have the same pressure ratio:

From the isentropic relations for an ideal gas, we know that pressure ratio, ratio, , are related by: .

, and temperature

The temperature ratios of any elementary cycle ``i'' are therefore the same and each of the elementary cycles has the same thermal efficiency. We only need to find the temperature ratio across any one of the cycles to find what the efficiency is. We know that the temperature ratio of the first elementary cycle is the ratio of compressor exit temperature to engine entry (atmospheric for an aircraft engine) temperature, in Figure 6.4. If the efficiency of all the elementary cycles has this value, the efficiency of the overall Brayton cycle (which is composed of the elementary cycles) must also have this value. Thus, as previously,

Figure 6.5: Arbitrary cycle operating between

A benefit of this view of efficiency is that it allows us a way to comment on the efficiency of any thermodynamic cycle. Consider the cycle shown in Figure 6.5, which operates between some maximum and minimum temperatures. We can break it up into small Carnot cycles and evaluate the efficiency of each. It can be seen that the efficiency of any of the small cycles drawn will be less than the efficiency of a Carnot cycle between and . This graphical argument shows that the efficiency of any other thermodynamic cycle operating between these maximum and minimum temperatures has an efficiency less than that of a Carnot cycle.

Muddy Points If there is an ideal efficiency for all cycles, is there a maximum work or maximum power for all cycles? (MP 6.7)

6.4.1 Net work per unit mass flow in a Brayton cycle


In Section 3.7.1 we found the net work of a Brayton cycle in terms of heat transfer. Now that we have defined entropy, we can reexamine the net work using an enthalpy-entropy ( - ) diagram, Figure 6.6. The net mechanical work of the cycle is given by:

where, by the first law,

If kinetic energy changes across the compressor and turbine are neglected, the temperature ratio, , across the compressor and turbine is related to the enthalpy changes:

The net work is thus

The turbine work is greater than the work needed to drive the compressor, as is evident on the ( ) diagram.

Figure 6.6: Brayton cycle in enthalpy-entropy (

- ) representation showing compressor and turbine work

Next: 6.5 Irreversibility, Entropy Changes, Up: 6. Applications of the Previous: 6.3 Representation of Thermodynamic Contents Index

UnifiedTP

Thermodynamics and Propulsion

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index Subsections

1.3.1 Heat 1.3.2 Zeroth Law of Thermodynamics 1.3.3 Work o 1.3.3.1 Example: Work on Two Simple Paths o 1.3.3.2 Example: Work Done During Expansion of a Gas

1.3.4 Work vs. Heat - which is which?

1.3 Changing the State of a System with Heat and Work


Changes in the state of a system are produced by interactions with the environment through heat and work, which are two different modes of energy transfer. During these interactions, equilibrium (a static or quasi-static process) is necessary for the equations that relate system properties to one-another to be valid.

1.3.1 Heat
Heat is energy transferred due to temperature differences only. 1. 2. 3. 4. Heat transfer can alter system states; Bodies don't ``contain'' heat; heat is identified as it comes across system boundaries; The amount of heat needed to go from one state to another is path dependent; Adiabatic processes are ones in which no heat is transferred.

1.3.2 Zeroth Law of Thermodynamics


With the material we have discussed so far, we are now in a position to describe the Zeroth Law. Like the other laws of thermodynamics we will see, the Zeroth Law is based on observation. We start with two such observations:
1. If two bodies are in contact through a thermally-conducting boundary for a sufficiently long time, no further observable changes take place; thermal equilibrium is said to prevail. 2. Two systems which are individually in thermal equilibrium with a third are in thermal equilibrium with each other; all three systems have the same value of the property called temperature.

These closely connected ideas of temperature and thermal equilibrium are expressed

formally in the ``Zeroth Law of Thermodynamics:''


Zeroth Law: There exists for every thermodynamic system in equilibrium a property called temperature. Equality of temperature is a necessary and sufficient condition for thermal equilibrium.

The Zeroth Law thus defines a property (temperature) and describes its behavior1.3. Note that this law is true regardless of how we measure the property temperature. (Other relationships we work with will typically require an absolute scale, so in these notes we use either the Kelvin or Rankine scales. Temperature scales will be discussed further in Section 6.2.) The zeroth law is depicted schematically in Figure 1.8.

Figure 1.8: The zeroth law schematically

1.3.3 Work
[VW, S & B: 4.1-4.6] Section 1.3.1 stated that heat is a way of changing the energy of a system by virtue of a temperature difference only. Any other means for changing the energy of a system is called work. We can have push-pull work (e.g. in a piston-cylinder, lifting a weight), electric and magnetic work (e.g. an electric motor), chemical work, surface tension work, elastic work, etc. In defining work, we focus on the effects that the system (e.g. an engine) has on its surroundings. Thus we define work as being positive when the system does work on the surroundings (energy leaves the system). If work is done on the system (energy added to the system), the work is negative. Consider a simple compressible substance, for example, a gas (the system), exerting a force on the surroundings via a piston, which moves through some distance, (Figure 1.9). The work doneon the surroundings, , is

therefore

Why is the pressure

instead of

? Consider

(vacuum). No work is done on the

surroundings even though

changes and the system volume changes.

Use of instead of is often inconvenient because it is usually the state of the system that we are interested in. The external pressure can only be related to the system pressure if . For this to occur, there cannot be any friction, and the process must also be slow enough so that pressure differences due to accelerations are not significant. In other words, we require a ``quasi-static'' process, . Consider .

Therefore, when

is small (the process is quasi-static),

and the work done by the system is the same as the work done on the surroundings.

Under these conditions, we say that the process is ``reversible.'' The conditions for reversibility are that:
1. If the process is reversed, the system and the surroundings will be returned to the original states. 2. To reverse the process we need to apply only an infinitesimal . A reversible process can be altered in direction by infinitesimal changes in the external conditions (see Van Ness, Chapter 2).

Remember this result, that we can only relate work done on surroundings to system pressure for quasi-static (or reversible) processes. In the case of a ``free expansion,'' where (vacuum), is not related to because the system is not in equilibrium. (and thus, not related to the work)

We can write the above expression for work done by the system in terms of the specific volume, ,

where is the mass of the system. Note that if the system volume expands against a force, work is done by the system. If the system volume contracts under a force, work is done on the system.

Figure 1.9: A closed system (dashed box) against a piston, which moves into the surroundings

Figure 1.10: Work during an irreversible process

For simple compressible substances in reversible processes, the work done can be represented as the area under a curve in a pressure-volume diagram, as in Figure 1.11(a).

[Work is area under curve of

[Work depends on

path] Figure 1.11: Work in coordinates

Key points to note are the following:


1. Properties only depend on states, but work is path dependent (depends on the path taken between states); therefore work is not a property, and not a state variable. 2. When we say , the work between states 1 and 2, we need to specify the path. ; either the work must be

3. For irreversible (non-reversible) processes, we cannot use given or it must be found by another method.

Muddy Points How do we know when work is done? (MP 1.3)

1.3.3.1 Example: Work on Two Simple Paths


Consider Figure 1.12, which shows a system undergoing quasi-static processes for which we can calculate work interactions as .

Figure 1.12: Simple processes

Along Path a:

Along Path b:

Practice Questions Given a piston filled with air, ice, a bunsen burner, and a stack of small weights, describe
1. how you would use these to move along either path a or path b above, and 2. how you would physically know the work is different along each path.

1.3.3.2 Example: Work Done During Expansion of a Gas


Consider the quasi-static, isothermal expansion of a thermally ideal gas from , to

, , as shown in Figure 1.13. To find the work we must know the path. Is it specified? Yes, the path is specified as isothermal.

Figure 1.13:Quasi-static, isothermal expansion of an ideal gas

The equation of state for a thermally ideal gas is

where is the number of moles, is the Universal gas constant, and is the total system volume. We write the work as above, substituting the ideal gas equation of state,

also for

, so the work done by the system is

or in terms of the specific volume and the system mass,

1.3.4 Work vs. Heat - which is which?


We can have one, the other, or both: it depends on what crosses the system boundary (and thus, on how we define our system). For example consider a resistor that is heating a volume of water (Figure 1.14):

Figure 1.14: A resistor heating water

1. If the water is the system, then the state of the system will be changed by heat transferred from the resistor. 2. If the system is the water and the resistor combined, then the state of the system will be changed by electrical work.

Next: 1.4 Muddiest Points on Up: 1. Introduction to Thermodynamics Previous: 1.2 Definitions and Fundamental Contents Index

UnifiedTP

Das könnte Ihnen auch gefallen