Fast n Furious

Fast n Furious
mechanical engineers can become a mechanic ,software engineers cannot become a software....

Dec 29, 2012

Automatic Pneumatic Bumper For Four Wheeler

Description: Many years ago, wheels were the part of a log and it slowly utilized for carts and wagons. The wooden wheel utilized was hard wood stakes. Trucks have become the backbone of the workforce in the world. They are large, strong and could be move on roughest of terrains. Truck rims should be placed if they are cracked.


The Automatic Pneumatic Bumper For Four Wheeler  of truck rims are manufactured in the similar manner. It begins with tough hub and 4 to 6 holes for the bolts. Truck wheels require durable which carry some weight. Lighter wheels are developed by decreasing unsprung mass and permit suspension to follow the terrain and develop grip. Better heat conduction spends heat from the brakes that develops braking function in driving situations and decreases the brake failure because of overheating. The spun steel rim is saved with welds series. The rim is balanced and provided the smooth finishing.
The wheel rim is the portion of automotive. It includes static loads and fatigue loads like wheel rim moves various road profile. It improves high stresses in rim to search the serious stress point. The alloy of steel and light are important materials utilized in a wheel consisting glass-fiber. Different wheels are Wire Spoke Wheel, Steel Disc Wheel, and Light Alloy Wheel. The steel rims are present in silver and chrome and many finishes are secured for alloy wheels.


The important components of the project are,
•  IR transmitter
•  IR receiver
•  Control Unit with Power supply
•  Solenoid Valve
•  Flow control Valve
•  Air Tank (Compressor)
The IR TRANSMITTER circuit is to transmite the Infra-Red rays. If any obstacle is there in a path, the Infra-Red rays reflected. This reflected Infra-Red rays are received by the receiver circuit is called " IR RECEIVER".


In this project, we have to apply this breaking arrangement in one wheel as a model. The compressed air drawn from the compressor in our project. The compressed air floe through the Polyurethene tube to the flow control valve. The flow control valve is connected to the solenoid valve as mentioned in the block diagram. The application of pneumatics produces smooth operation. By using more techniques, they can be modified and developed according to the applications


Conclusion:

The wheel rim is the outer circular concept of the metal where corner of the tyre is climbed on automobiles like vehicles. The automotive steel wheel rim is created with rectangular metal. The metal plate is made curved to generate the cylindrical sleeve with the two corners of the sleeve joined together.

Solar Sail


It's interesting when you look back at the history of space exploration and realize that propulsion technology hasn't really changed very much.
The earliest rocket prototypes were nothing more than elaborate versions of weapons used during World War 2 and fireworks used during civil celebrations. Even the Space Shuttle made use of solid rocket fuel technology in its pair of solid rocket boosters. But, with the liquid rocket fuel propulsion in the external tank, this combination has proved to be highly effective and launched hundreds of astronauts into space.
The approach works -- albeit not very efficiently -- and to get out of the Earth's gravitational well, it seems for now that the extra punch from exothermic processes is needed.
In deep space, however, there are alternatives receiving very serious consideration -- such as the "eco-friendly" solar sail.

What is a Solar Sail?
A solar sail, simply put, is a spacecraft propelled by sunlight. Whereas a conventional rocket is propelled by the thrust produced by its internal engine burn, a solar sail is pushed forward simply by light from the Sun. This is possible because light is made up of packets of energy known as “photons,” that act like atomic particles, but with more energy. When a beam of light is pointed at a bright mirror-like surface, its photons reflect right back, just like a ball bouncing off a wall. In the process the photons transmit their momentum to the surface twice – once by the initial impact, and again by reflecting back from it. Ever so slightly, propelled by a steady stream of reflecting photons, the bright surface is pushed forward.
A solar sail is made up of just such a reflective surface, or several surfaces, depending on the sail’s design. When the bright sails face the Sun directly, they are subjected to a steady barrage of photons that reflect off the shiny surfaces and impel the spacecraft forward, away from the Sun. By changing the angle of the sail relative the Sun it is possible to affect the direction in which the sail is propelled – just as a sailboat changes the angle of its sails to affect its course. It is even possible to direct the spacecraft towards the Sun, rather than away from it, by using the photon’s pressure on the sails to slow down the spacecraft’s speed and bring its orbit closer to the Sun.
In order for sunlight to provide sufficient pressure to propel a spacecraft forward, a solar sail must capture as much Sunlight as possible. This means that the surface of the sail must be big – very big. Cosmos 1 is a small solar sail intended only for a short mission. Nevertheless, once it spreads its sails even this small spacecraft will be 10 stories tall, as high as the rocket that will launch it. Its eight triangular blades are 15 meters (49 feet) in length, and have a total surface area of 600 square meters (6500 square feet). This is about one and a half times the size of a basketball court.
For a true exploration mission the requirements are far greater: when a NASA team in the 1970s, headed by Louis Friedman, suggested using a solar sail spacecraft for a rendezvous with Halley’s comet, they proposed a sail with a surface area of 600,000 square meters (6.5 million square feet). This is equivalent to a square of 800 meters (half-mile) by 800 meter – the size of 10 square blocks in New York City!
Even with such a gigantic surface, a solar sail spacecraft will accelerate very slowly when compared to a conventional rocket. Under optimal conditions, a solar sail on an interplanetary mission would gain only 1 millimeter per second in speed every second it is pushed along by Solar radiation. The Mars Exploration Rovers, by comparison, accelerated by as much as 59 meters (192 feet) per second every second during their launch by conventional Delta II rockets. This acceleration is 59,000 times greater than that of a solar sail!
But the incomparable advantage of a solar sail is that it accelerates CONSTANTLY. A rocket only burns for a few minutes, before releasing its payload and letting it cruise at a constant speed the rest of the way. A solar sail, in contrast, keeps on accelerating, and can ultimately reach speeds much greater than those of a rocket-launched craft. At an acceleration rate of 1 millimeter per second per second (20 times greater than the expected acceleration for Cosmos 1), a solar sail would increase its speed by approximately 310 kilometers per hour (195 mph) after one day, moving 7500 kilometers (4700 miles) in the process. After 12 days it will have increased its speed 3700 kilometers per hour (2300 mph).
While these speeds and distances are already substantial for interplanetary travel, they are insignificant when compared to the requirements of a journey to the stars. Given time, however, with small but constant acceleration, a solar sail spacecraft can reach any desired speed. If the acceleration diminishes due to an increasing distance from the Sun, some scientists have proposed pointing powerful laser beams at the spacecraft to propel it forward. Although such a strategy is not practicable with current technology and resources, solar sailing is nevertheless the only known technology that could someday be used for interstellar travel.


 How do they work?
Solar Sails propel a spacecraft by utilizing the pressure created by the stream of photons (tiny units of light energy) from the sun. Once a spacecraft is in orbit, a lightweight sail would unfurl. Changing the position of the sail would increase or decrease speed. The thrust created by the photon stream is very low and interplanetary journeys would take years. For long missions, an on-board laser or microwave transmitter would be fitted to provide power when the Sun is distant.


Hundreds of space missions have been launched since the last lunar mission, including several deep space probes that have been sent to the edges of our solar system. However, our journeys to space have been limited by the power of chemical rocket engines and the amount of rocket fuel that a spacecraft can carry. Today, the weight of a space shuttle at launch is approximately 95 percent fuel. What could we accomplish if we could reduce our need for so much fuel and the tanks that hold it?
International space agencies and some private corporations have proposed many methods of transportation that would allow us to go farther, but a manned space mission has yet to go beyond the moon. The most realistic of these space transportation options calls for the elimination of both rocket fuel and rocket engines -- replacing them with sails. Yes, that's right, sails.
NASA is one of the organizations that has been studying this amazing technology called solar sails that will use the sun's power to send us into deep space.


History
On February 4, 1993, Znamya 2, a 20-meter wide aluminized-mylar reflector, was successfully tested from the Russian Mir space station. Although the deployment test was successful, the experiment only demonstrated the deployment, not propulsion. A second test, Znamaya 2.5, failed to deploy properly.
On August 9, 2004, the Japanese ISAS successfully deployed two prototype solar sails from a sounding rocket. A clover type sail was deployed at 122 km altitude and a fan type sail was deployed at 169 km altitude. Both sails used 7.5 micrometer thick film. The experiment was purely a test of the deployment mechanisms, not of propulsion.
In June 21, 2005 a Volna rocket launched from a Russian submarine in the Barents Sea launched the privately built Cosmos-1 spacecraft. However, a rocket failure prevented it from reaching its intended orbit. The Cosmos-1 spacecraft was designed to use solar sails to move through space. Had the mission been successful, it would have been the first ever orbital use of a solar sail to speed up a spacecraft, as well as the first space mission by a space advocacy group (Planetary Society).
The IKAROS probe is the world's first spacecraft to use solar sailing as the main propulsion.
LightSail-1, a second orbital spacecraft by the Planetary Socity is under construction and is expected to be ready by the end of 2010.


How It Works: The sun’s radiation exerts force on the ultrathin fabric of solar sails, much like wind propels a sailboat. These sails can clean up orbit by slowing debris enough that it deorbits. For example, an expandable 97-square-foot sail can be launched as a secondary payload on even a small 110-pound satellite, and an onboard system or ground control triggers its deployment. Conductive coils embedded in the sail control its angle, so it can maneuver a satellite out of orbit; sail and satellite disintegrate together in the atmosphere.
Pros: The material is cheap and portable.
Cons: The sail’s expansion and the spacecraft’s altitude need to be carefully calculated beforehand.
Plausibility: The technology has existed for decades, from a sail-equipped craft in the mid-1970s intended to ride along with Halley’s Comet (the craft never flew, and the project was scrapped) to large solar arrays currently affixed to the Messenger spacecraft, which are helping steer it to Mercury. Most recently, the Japan Aerospace Exploration Agency launched a solar-sail-propelled “space yacht” called Ikaros in May. But the solar sails precisely controllable enough to remove debris are still years away.
Cloud Cover: Tanks of liquid gas would emit mist that slows down objects passing through enough that the debris would fall from orbit.  Kevin Hand

Space Mist

How It Works: One of the more novel solutions, first proposed by researchers at NASA’s Ames Research Center in 1990 and recently resurrected, is to use frozen mist to drag an object out of orbit. A rocket launches a tank filled with liquid gas, such as carbon dioxide, and thrusters position the tank near the path of the target object.Thousands of miles from the debris, the tank sprays a cloud of frozen mist. The droplets slow down and deorbit anything they encounter. According to the idea’s originator, aerospace engineer George Sarver of Ames, a spaceship emitting a 220-pound, 62-foot-diameter cloud of frozen mist could deorbit dense objects such as steel nuts. Larger clouds could stop less-dense items like insulation.
Pros: Once the mist dissipates, nothing remains in orbit. The tank falls back into the atmosphere, which means less clutter that could otherwise create more debris.
Cons: It requires precise aim, because each tank sprays its shot just once.
Plausibility: “We could probably test it in orbit within 18 months if NASA funded it,” says Creon Levit, the chief scientist for programs at Ames.
Astro-Velcro: Large balls with adhesive shells could stick themselves to debris and carry it out of orbit. Kevin Hand

Robots and Adhesives

How It Works: One idea to capture tumbling debris employs orbiting spacecraft that would use robotic arms to grab and release debris, essentially tossing it out of orbit so it burns up in the atmosphere. Sean Shepherd, a curriculum coordinator at Eastern New Mexico University, proposes a more novel idea, called Adhesive Synthetic Trash Recovery Orbital Spheres, or Astros—essentially a collection of sticky balls. The balls consist of a layer of metallic foam (such as silicon carbide) and an outer shell of adhesive (such as aerogels or resins). They attach themselves to debris and then dive into the atmosphere, where both are incinerated.
Pros: Spacecraft that carry and position the robotics are easier to control than tethers or sails and can remove multiple pieces of debris. Adhesives are cheap.
Cons: The adhesive balls can’t be compressed, so packing them into a rocket for launch will be difficult.
Plausibility: Hanspeter Schaub, the associate chair of graduate studies in the Aerospace Engineering Sciences department at the University of Colorado, says the orbiting robot concept could be tested in space within five years.



Dec 21, 2012

Hybrid hydraulic Vehicle



Introduction To Hydraulic Hybrid Vehicles:
Hybrid vehicles use two sources of power to drive the wheels. In a hydraulic hybrid vehicle (HHV) a regular internal combustion engine and a hydraulic motor are used to power the wheels.
Hydraulic hybrid systems consist of two key components:
  • High pressure hydraulic fluid vessels called accumulators, and
  • Hydraulic drive pump/motors.
Working of Hydraulic Hybrid Systems:
01-hydraulic-hybrid-retrofit-hydraulic hybrid system-HHS-regenerating braking energy
The accumulators are used to store pressurized fluid. Acting as a motor, the hydraulic drive uses the pressurized fluid (Above 3000 psi) to rotate the wheels. Acting as a pump, the hydraulic drive is used to re-pressurize hydraulic fluid by using the vehicle’s momentum, thereby converting kinetic energy into potential energy. This process of converting kinetic energy from momentum and storing it is called regenerative braking.
The hydraulic system offers great advantages for vehicles operating in stop and go conditions because the system can capture large amounts of energy when the brakes are applied.
The hydraulic components work in conjunction with the primary. Making up the main hydraulic components are two hydraulic accumulator vessels which store hydraulic fluid compressing inert nitrogen gas and one or more hydraulic pump/motor units.
The hydraulic hybrid system is made up of four components.
  • The working fluid
  • The reservoir
  • The pump or motor
  • The accumulator
The pump or motor installed in the system extracts kinetic energy during braking. This in turn pumps the working fluid from the reservoir to the accumulator, which eventually gets pressurized. The pressurized working fluid then provides energy to the pump or motor to power the vehicle when it accelerates. There are two types of hydraulic hybrid systems – the parallel hydraulic hybrid system and the series hydraulic hybrid system. In the parallel hydraulic hybrid, the pump is connected to the drive-shafts through a transmission box, while in series hydraulic hybrid, the pump is directly connected to the drive-shaft.
There are two types of HHVs:
  • Parallel and
  • Series.
Parallel Hydraulic Hybrid Vehicles:
01-hydraulic hybrid cars-HLA system-pump mode to motor mode-parallel hydraulic hybrid vehicles-nitrogen accumulator pressure 5000 psi
In parallel HHVs both the engine and the hydraulic drive system are mechanically coupled to the wheels. The hydraulic pump-motor is then integrated into the driveshaft or differential.
Series Hydraulic Hybrid vehicles:
01-hydraulic hybrid vehicles-combines regular internal combustion engine- hydraulic motor as a accumulator-kinetic energy into potential energy to drive the vehicle
Series HHVs rely entirely on hydraulic pressure to drive the wheels, which means the engine does not directly provide mechanical power to the wheels. In a series HHV configuration, an engine is attached to a hydraulic engine pump to provide additional fluid pressure to the drive pump/motor when needed.
Advantages:
  • Higher fuel efficiency.  (25-45 percent improvement in fuel economy)
  • Lower emissions.  (20 to 30 percent)
  • Reduced operating costs. 
  • Better acceleration performance.

Jul 18, 2012

How to build a Solar Car



Building a solar car is a very ambitious goal, but one that is achievable through hard work and commitment. One of the first things to do is to organize a group of people interested in building a car, and planning a specific goal. How much money will you need to build a solar car? If the least expensive parts are used, the cost of a solar car is about $12,000, not including trip costs to the competition. You will need a core group of fundraisers who are willing to spend the year making presentations to businesses and speaking to individuals at their home. You will also need to assess what materials and tools you need and get several people focused on building the car itself. It may be a difficult road the first year, but the things you learn along the way will last a lifetime!
Solar cars are powered by the sun's energy. The main component of a solar car is its solar array, which collect the energy from the sun and converts it into usable electrical energy. The solar cells collect a portion of the sun's energy and stores it into the batteries of the solar car. Before that happens, power trackers converts the energy collected from the solar array to the proper system voltage, so that the batteries and the motor can use it. After the energy is stored in the batteries, it is available for use by the motor & motor controller to drive the car. The motor controller adjusts the amount of energy that flows to the motor to correspond to the throttle. The motor uses that energy to drive the wheels.

Solar Array and Power Trackers
We recommend a solar array created from individual solar cells as opposed to one made of prefabricated solar panels. It enhances the students' learning and can result in a lighter solar array. Cells can be bought from either Siemens or ASE Americas. Both sell the terrestrial-grade cells that are permitted in the Winston Solar Challenge, and the cost for terrestrial-grade cells are much lower than space-grade cells, though terrestrial-grade is less efficient. Each solar cell should produce .5 volts at about 3 amps at peak sunlight. The number of cells to use depends on their size and the allowable solar area per Winston rules. Solar cells should be wired in series on a panel and should be divided into several zones. For example, if you have 750 solar cells, you might want to wire 3 sets of 250 cells, each zone producing about 125 volts. If one zone fails, two other zones are still producing power. The solar array voltage does not need to match the system voltage of the motor if you use power trackers. Power trackers convert the solar array voltage to the system voltage. They are essential in a solar car. Be sure to verify with the power tracker vendor the necessary array voltage to feed the power trackers. If the car drives underneath shade, the power trackers automatically adjusts the power to match system voltage, allowing the system to run as efficient as possible. Power trackers are available from AERL.
Batteries
The batteries store energy from the solar array and makes them available for the motor's use. Many different types of batteries are sold. Most high school teams use lead-acid batteries because they are inexpensive, but some teams use lithium-ion or nickel-cadmium. We recommend that you stick with lead-acid batteries because they are readily available and inexpensive. Another choice teams must make is running with flooded-cell batteries or gel-cell batteries. Flooded-cell batteries are the standard automotive batteries filled with liquid sulfuric acid. They are preferred because they can be overcharged without risk of blowing up, but they weigh more than gel-cell batteries. Gel-cell are sealed and lightweight, but when charging the batteries, check the battery voltage often. The number of batteries to choose depends on the motor (system) voltage. If the system voltage is 72 volts, you will need 6 12-volt batteries. Also be sure to check the rules for weight or watthour requirements. Buy batteries with as many amphours as allowed by the rules to maximize the amount of energy you can store.
Motor & Controller
Most teams use DC brush permanent magnet motors to drive their solar cars. Inexpensive and easy to hook up, these motors are desirable for high school teams with little financial support. Expect a maximum efficiency of 80-90%. For teams with more money, brushless motors increase the efficiency of the motor to the 94-99% range. Also, some motor and controller setups allow for regenerative braking, which allows the solar car to put energy back into the batteries when going downhill. For the beginning team, DC brush motors would be sufficient to get a solar car up and running. Another variable in choosing a motor is how much power it has. We have found that there is little need to have more than 5hp continuous power output on our motors. There are two manufacturers who supply most teams with motors and controllers: Solectria and Advanced DC Motors. Many college teams buy their motors from Solectria, but Advanced DC Motors have less expensive motors. Controllers usually drive a particular motor. Once you choose the motor that suits your needs, the same vendor would most likely have a matching controller.
Instrumentation
One of the most important pieces of instrumentation is a state-of-charge meter. A state-of-charge meter gives information about system voltage, amp draw, battery energy remaining, and estimates the how much time remains until the battery is out of energy. We found that the E-Meter, manufactured by Cruising Equipment, served out purpose well. It has a digital display and accurately counts the number of amp-hours remaining in the battery. The E-Meter is the do-it-all in instrumentation. Another instrument that may be useful is a speedometer. Instead of using a regular speedometer drive, use magnetic contact speedometers, found in many sports equipment stores. This option does not add drag to your car. To ensure that your batteries are running properly, you may invest in getting a voltmeter for each of your batteries. A failed battery may show the proper voltage when the car is not running, but while the battery is under load, the voltmeter will show a lower than normal battery voltage.
Steering & Suspension
We strongly recommend front wheel steering as it tends to be more stable and safer. A solar car uses energy frugally if it is to be competitive. If there are two front wheels, it is therefore advisable to work out the geometry so that they run parallel when the car is going straight ahead, but when the car is turning, the front wheels turn at different radii. If the car is turning left, the left front tire is making a smaller circle than the right front tire. If the tires remain parallel while turning, they will cause unnecessary drag, decreasing tire life and overall performance.
The only advice we can offer with respect to suspension is that it should be soft enough to protect the car and solar array from unnecessary jolts and firm enough to provide a stable ride.
Brakes
Disc brakes are desirable as they are predominantly hydraulic. Having hydraulic lines running to the wheels can be easier than mechanical brake arrangements. The most significant problem with disc brakes is that the brake pads do not back away from the brake rotors when pressure is released, they just relieve braking pressure. Because the pads don't normally back away from the rotors, they continue to have a small amount of drag. While this drag may not be noticeable on the family car, it is very inefficient on solar cars. Go kart shops now have brake calipers that are spring loaded to move the pads away from the rotors. We have found these very worthwhile.
Tires & Hubs
Tire selection will affect rolling resistance which affects how far the solar car will travel with the energy available. Tires with thicker rubber and wider tread tend to have higher rolling resistance (a bad thing). Thinner tires with higher pressuer have less rolling resistance, but are more susceptible to flats. The best tires we have found are the Bridgestone Ecopia tires made for solar cars. They are very thin and operate at over one hundred pounds/inch pressure. Unfortunately, they need to be mounted on specially made wheels and require custom made hubs. On the good side, these tires and wheels are very light. Some colege teams have experimented with bicycle tires but report limited success (bicycle tires, rims and spokes are not designed for the forces placed on them by non-tilting vechicles that weigh several hundred pounds). Motorcycle tires tend to have more resistance, although there may be high pressure tires with low resistance that we don't know about yet.
Bearing resistance can be reduced by light minimal lubrication. Bearing seals can be cut away at the contact lip to leave most of the seal protection while removing most if not all seal drag. It is a good idea to get the rolling chassis operational months before your schedule gets critical. Run the chassis as many miles as possible to prove that your bearings, axles, steering and suspension can survive.

Jun 28, 2012

Thermal Cycles




Otto Cycle
1-2             Reversible adiabatic compression of air (WIN)
2-3             Addition of heat at constant volume increases pressure (QIN)
3-4     Reversible adiabatic expansion of air (WOUT)
4-1     Rejection of heat at constant volume reduces pressure (QOUT)

§         From first law 
§         Thermal efficiency of a cycle usually found .  Also  and  hence 
§         Compression and expansion of air are Polytropic process with  hence  Where the compression ratio 
§         Hence thermal efficiency is only dependant on compression ratio 


Diesel cycle
1-2             Reversible adiabatic compression of air (WIN)
2-3             Addition of heat at constant pressure increases volume (QIN)
3-4     Reversible adiabatic expansion of air (WOUT)
4-1     Rejection of heat at constant volume reduces pressure (QOUT)

§         From first law 
§         As with Otto  but as heat is added at constant pressure 
§         Assume heat addition 2-3 is ideal gas:  but  so  hence the cut-off ratio 
§         Thermal efficiency is dependant on compression AND cut-off ratio 




Real gas turbines
§         Approximated by joule cycle. 
§         Air drawn into compressor.
§         Fuel injected and ignited.
§         Hot gasses pass through turbine, providing mechanical power.
§         Exhaust gasses expelled.






Joule Cycle
§         Approximates gas turbine, like aircraft engines.
§        
Open system so use SFEE!


1-2             Reversible adiabatic compression of air (WIN)
2-3             Reversible addition of heat at constant p increases T (QIN)
3-4     Reversible adiabatic expansion of air (WOUT)
4-1     Reversible rejection of heat at constant p reduces T (QOUT)

§         From first law 
§         For heat transfers at constant pressure  and 
§         For compressor and turbine work  and 
§         Compression and expansion is Polytropic with  such that 
§         Here we have the pressure ratio 
§         Thermal efficiency dependant only on the pressure ratio 


Otto cycle
The Otto cycle is a first approximation to model the operation of a spark-ignition engine, first built by Nikolaus Otto in 1876, and used in many cars, small planes and small power systems (below say 200 kW) down to miniature engines. This is a reciprocating internal combustion gas engine, in contrast to the, at that time master, reciprocating or rotodynamic external combustion steam engine. The Otto engine is sketched in Fig. 5.2, where the typical terms are introduced for engine-geometry characteristics (stroke, bore, displacement and compression ratio); other terms for engine-operation characteristics are shaft speed, mean pressure, power, fuel consumption, torque, volumetric efficiency, energy efficiency, etc.

Fig. 5.2. Sketch and nomenclature for reciprocating engines: a) 4-stroke, b) 2-stroke (uni-flow).

In the ideal air-standard Otto cycle, the working fluid is just air, which is assumed to follow four processes (Fig. 5.3): isentropic compression, constant-volume heat input from the hot source, isentropic expansion, and constant-volume heat rejection to the environment.

Fig. 5.3. The ideal Otto cycle in the T-s and p-V diagram, and a practical p-V trace of four-stroke and two-stroke engines.

The main parameters of both ideal and real Otto cycles are:
·        Size, measured by the displacement volume (the volume swept by the piston, V1V2), usually less than 0.5 litres per cylinder, to avoid self-ignition.
·        Speed, more precisely crankshaft speed, n, with a typical operation range n=1000..7000 rpm (n=20..120 Hz). The maximum value may be in the range nmax=6000..8000rpm for four-stroke engines. Two-stroke motorcycle engines run quicker (nmax=13 000 rpm), the quickest (nmax=20 000 rpm) being the smallest engines (two stroke), used in aircraft modelling.
·        Compression ratio, r=V1/V2, with a typical range of r=8..10 (up to 14 in direct-injection spark-ignition engines), limited by the 'knock' or self-ignition problem.
·        Mean effective pressure, pme, defined as the unit work divided by the displacement, with a typical range of 0.2..1.5 MPa (the full-load value may range from pme=1.2 MPa in two-stroke motorcycle engines, to pme=1.7 MPa in the largest turbocharged engines). Maximum pressure may have a typical range of 4..10 MPa. Performance maps of reciprocating engines are usually presented on a pme-n diagram, i.e. mean effective pressure versus engine speed.

Real reciprocating engines may be two-strokes or four-strokes in a cycle; the simple to model is the four-strokes engine, where one stroke is used to fill the cylinder with the air-fuel mixture through the inlet valve (the inside pressure is a little below atmospheric), a second stroke to compress the mixture with valves closed and exploding it at a precise point with a spark, a third stroke to let the product gasses to expand and do work on the crank-shaft, and a fourth stroke to force the burnt gasses out through the exhaust valve, and start a new cycle (two crank-shaft turns have elapsed). In two-stroke engines there is an overlap between intake and exhaust, with the inlet stream used to sweep the burnt gasses (throwing unburnt gases to the tailpipe!), and the whole cycle is performed in one turn of the crankshaft. In spite of the higher fuel consumption and pollutant emissions, and the difficulties for their lubrication (oil is added to the fuel), two-stroke engines have very high specific power, and thus they are used for small engines (hand-held and garden engines, motorcycles and outboard boats).

The cold-air-standard model takes as working fluid air with constant properties (those at the inlet, i.e. cold), what renders the analysis simple. The energy exchanges for the trapped control mass, m, are W12/m=cv(T2-T1), Q12=0, W23=0, Q23/m=cv(T3-T2), W34/m=cv(T4-T3), Q34=0, W41=0, Q41/m=cv(T1-T4), and the energy and exergy efficiencies are:

                     and    (5.2)

Real spark-ignition engines do not recycle the working fluid; they are all internal combustion engines where ambient air is pumped in, and petrol fuel (gasoline) is added  to prepare a homogeneous reactive mixture that at a certain point (cinematically or electronically controlled) is ignited by an electrical spark. Fuel addition can be in a carburettor by Venturi suction, or in the individual inlet duct by injection, or recently by direct injection inside the cylinder as in a diesel engine (but with a much lower injection pressure, less than 10 MPa). Gaseous fuels like coal-gas, liquefied petroleum gas (LPG), and natural gas are also used, as well as other liquid fuels like methanol, ethanol and ethers, usually added to gasoline up to a 20%. Typical energy efficiencies are low, 25% to 35% when running at nominal power, much lower at partial load, but the engine is light, very powerful and not expensive.

Diesel cycle
The Diesel cycle is a first approximation to model the operation of a compression-ignition engine, first built by Rudolf Diesel in 1893, and used in most cars, nearly all trucks, nearly all boats, many locomotives, some small airplanes, and many large electric power systems and cogeneration systems. It is the reference engine from 50 kW to 50 MW, due to the fuel used (cheaper and safer than gasoline) and the higher efficiency. This is a reciprocating internal combustion engine with the same sketch as in Fig. 5.2, but substituting the spark-plug by the fuel injector (the fuel is injected at very high pressures, up to 200 MPa, to ensure immediate vaporisation). One of its key advantages compared to Otto engines is the great load increase per cylinder associated to the higher pressures allowed (the mixture of fuel and air would detonate in Otto engines at high compressions), and the further load increase associated to charging previously-compressed air (turbocharging). Another advantage is the much better performance at part load, since there it is achieved by injecting less fuel instead of by throttling, and the torque (power divided per angular speed, M=P/w), changes less with angular speed.

In the ideal air-standard Diesel cycle, the working fluid is just air, which is assumed to follow four processes (Fig. 5.4): isentropic compression, constant-pressure heat input from the hot source, isentropic expansion, and constant-volume heat rejection to the environment.

Fig. 5.4. The ideal Diesel cycle. Practical p-V diagrams are as in Fig. 5.3.

Similarly to the Otto cycle, the main parameters of ideal and real Diesel cycles are also the size, measured by the displacement volume (that may reach more than 1 m3 per cylinder in large marine engines), the compression ratio, r=V1/V2 (with a typical range of 16..22, limited just by strength), the cut-off ratio, rc, or the mean effective pressure (in the range 1..2 MPa), or the maximum pressure (in the range of 3 MPa to 20 MPa), and the speed (with a typical range of 100..6000 rpm). The energy efficiency can be expressed as:

                                                                                                                     (5.3)
                  
Real compression-ignition engines take ambient air (often after a first stage compression) and compress it (inside the cylinder) so much, rising the temperature accordingly, that the fuel burns as it is injected (after a small initial delay due to vaporisation and combustion kinetics). The external compression is performed in a centrifugal compressor driven by a centrifugal turbine moved by the exhaust gasses (turbocharger); the air is cooled after external compression (inter-cooler) before further compression within the cylinders, to increase the efficiency, as explained in Multistage compression.

The higher pressures in Diesel than in Otto engines require a robust engine-frame and delicate fluid injection hydraulics (with injection pressure up to 200 MPa), but the wider range of fuels (from gas-oils to fuel-oils), the better fuel control, longer durability and economy are making Diesel engines to displace traditional gasoline-engines markets (for heavy-duty applications it has always been unrivalled). Typical energy efficiencies are from 30% to 54% (based on the lower heating value of the fuel), the latter, the largest of any single thermal engine, being achieved in large two-stroke marine low-speed engines with bores larger than 0.5 m: first, because the thermal losses decrease with size, and second, because the very low speed (some 100 rpm instead of the typical 3000 rpm for a car engine) allows for a more complete combustion (more time to burn, and burning nearly without volume change) and decrease friction losses (in spite of the fact that the mean piston speed stays at some 6..7 m/s for the whole range of reciprocating engines: from the 1 cm3 50 W model, to a 1 m3 5 MW 'three-store castle', per cylinder). Most large Diesel engines are supercharged, i.e. fed with compressed air instead of atmospheric air, usually by means of a turbocharger (a small compressor shaft-coupled to a small turbine driven at high speeds by the exhaust gases; some 10% of the fuel energy goes through that shaft), with an intermediate cooling of the compressed air before intake to the cylinders (intercooler). The two-stroke cycle is better suited to Diesel engines, since only air is used to sweep the burnt gases (scavenging), and not fresh mixture, but, because of the difficulty in lubricating, it is only used in the largest marine engines (10 MW..100 MW), where residual fuel (must be preheated to flow) can be used.


Mixed cycle
The mixed (or dual, or Sabaté, or Seiliger, or 5-point) cycle, sketched in Fig. 5.5, is a refinement to both Otto and Diesel cycles, at the expense of an additional parameter, the heat-addition pressure ratio, rp=p3/p2.
  
Fig. 5.5. The dual or Sabaté cycle.

The energy efficiency can be expressed as:

                                                                                                         (5.4)


Brayton cycle
The Brayton cycle, named after the American engineer George Brayton (that built a two-stroke reciprocating engine in 1876 and advanced combustion chambers at constant pressure), is a good model for the operation of a gas-turbine engine (first successfully tested by F. Whittle in 1937, and first applied by the Heinkel Aircraft Company in 1939), nowadays used by practically all aircraft except smaller ones, by many fast boats, and increasingly been used for stationary power generation, particularly when both power and heat are of interest.

In the ideal air-standard Brayton cycle, the working fluid is just air, which is assumed to follow four processes (Fig. 5.6): isentropic compression, constant-pressure heat input from the hot source, isentropic expansion, and constant-pressure heat rejection to the environment. Contrary to reciprocating engines, the gas turbine is a rotatory device working at a nominal steady state (it can hardly work at partial loads); spark ignition is used to start up, since air compressor output temperature is not high enough to inflame the fuel).

Fig. 5.6. The ideal Brayton cycle in the T-s and p-V diagram, and the regenerative Brayton cycle.

Most gas turbines are internal combustion engines where the working fluid must be renovated continuously as sketched in Fig. 5.7, but some gas turbines use a closed-loop working fluid.

Fig. 5.7. Open and closed cycle gas turbine.

The main parameters of ideal and real Brayton cycles are the turbine-inlet temperature, T3, the compressor pressure ratio, p=p2/p1 (with a typical range of 4 to 30), the compressor and turbine efficiencies, and the size, measured by the air mass flow rate. The energy and exergy efficiencies for the ideal Brayton cycle (compressor and turbine efficiencies of 100%), can be expressed as:

                      and                                 (5.5)

Real engine efficiencies are comparatively low, from 25% to 38%, but in combination with a bottoming vapour cycle, they reach 50..59% based on LHV; combined cycle power plants are the present standard in electricity generation. Contrary to reciprocating and steam engines, the gas turbine can only work with fine-tuned components, since it gives no net power if the compressor and turbine efficiencies fall below say 80% (modern gas turbines can have compressor efficiencies of 68% to 88%, and turbine efficiencies of 88% to 90%). It also needs high turbine-inlet temperature (modern aircraft gas turbines with blade cooling operate at up to 1700 K and at pressure ratios up to 30:1). It is easy to prove that for fixed extreme temperatures (ambient and turbine-inlet) there is a pressure ratio that maximises the work per unit mass flow rate, thus rendering the smallest engine for a given power, this optimum value being:

                                                                                                                        (5.6)

where hC and hT are the compressor and turbine isentropic efficiencies. .

Several improvements to the simple Brayton cycle are in use. Besides the multistage compression and expansion, the main variant is the regenerative cycle (Fig. 5.6), where heat from the exhaust gasses is used (from point 4 to 5) to heat up air before entering the combustion chamber (from point 2 to ideally up to point 2' in Fug. 5.6, although in practice the heat exchanger efficiency will limit this value). The heat recovery from the exhaust gasses may be also performed externally to the cycle, e.g. generating vapour in a heat exchanger (boiler), that may be directly used for heating applications or may even get expanded in a vapour turbine to produce further work (combined Brayton and Rankine cycles).


Stirling cycle and other gas cycles
Practical gas engines reduce to reciprocating engines and gas turbines, whose processes best represented by the dual cycle and the Brayton cycle, respectively, but there are more cycles of academic interest, as the Stirling cycle that models Stirling engines, invented in 1816 by the Scottish clergyman Robert Stirling, that, although faded away in the XIX c. against the steam and later gas engines, has got considerable attention lately.

The Stirling engine is an external combustion machine (all petrol and diesel engines, and practically all gas turbine engines, are of the internal combustion type, although, as said in the study of their cycles, they could operate as external combustion machines, in principle). The working substance may be just air or better helium, and the four processes followed are (Fig. 5.8): isothermal compression with heat input from the hot source, constant-volume compression by heat input from a regenerator (a porous solid matrix; an auxiliary displacer forces the gas through), isothermal expansion with heat rejection to the environment, and constant-volume heat rejection to the regenerator. Similar to the Stirling cycle is the Ericson cycle, where heat regeneration is isobaric instead of isometric as in the Stirling one (the regenerative Brayton cycle with infinitesimal multistage compressions and expansions approaches to the Ericson cycle). Notice that there is nothing essential to the four-process engines, and a three-process cycle (Fig. 5.8, named after J.J.E. Lenoir, a pioneer of two-strokes engines in the XIX c.) is used to model the operation of a pulse-reactor.

Fig. 5.8 The ideal Stirling cycle in the T-s and p-V diagram, and the ideal Ericson and Lenoir cycles.

Energy and exergy efficiencies of 36% and 50% respectively have been reached with prototype Stirling engines of up to 10 kW, the main problems being the regenerator loss of efficiency at high speeds (>30 Hz, i.e. >1800 rpm), the radiant heat looses at high temperature (>1000 K), incomplete exchange of gas between the hot and cold zones, and leaking at high pressure (>5 MPa).

Vapour power cycles

Rankine cycle
Most large electricity generating plants (central power stations), and very large ship engines, use water vapour (steam) as working fluid, following some variation of the basic Rankine cycle (named after the Scottish inventor William Rankine, that in 1859 wrote the first book on Thermodynamics), the only vapour power cycle in practical use since 1840 until in 1984 Alexander Kalina patented in the USA the cycle named after him.

The heat source for the boiler is usually the combustion products of a fuel (mainly coal) and air, or the primary refrigerant of a nuclear reactor, and the heat sink in the condenser is usually a water loop, open like in a river, or closed like in a cooling tower (as explained in Chapter 8). Thomas Newcomen is credited with the invention of the steam engine in 1705 for the purpose of driving the pumps used in clearing groundwater from mine shafts. Although the work-producing element was initially reciprocating cylinder-piston devices, in 1882 Gustav de Laval introduced the vapour turbine that has taken over.


Fig. 5.9. Carnot cycle within the two-phase region, basic Rankine cycle in the T-s and p-V diagram, and sketch of a vapour plant.

The four processes in a simple Rankine cycle are: isentropic compression of the liquid from 1 to 2 (Fig. 5.9), isobaric heating of the liquid, boiling and super-heating the vapour (from 2 to 3), isentropic expansion from 3 to 4, and isobaric heat rejection until full condensation of the vapour. The Rankine cycle is less efficient than the Carnot cycle (Fig. 5.9), but it is more practical since the compression is not in the two-phase region and only requires a small work, and the expansion is mainly in the gaseous phase (high-speed droplets erode turbine blades). Water is not the ideal working substance because it changes phase at relatively low temperatures (below the critical point at 647 K), generating a lot of entropy in the heat transfer from typical high-temperatures heat-sources: 1000 K in nuclear reactors up to 2000 K in conventional combustion plants. Nevertheless, water is practically the only working substance used, because of its good thermal properties and availability. A caution note is that the right-hand-side end of the two-phase region, the saturated vapour line, that for water in the T-s diagram has the shape shown in Fig. 5.9, may be more vertical and even have negative slope at low temperatures for heavier molecular substances, naturally avoiding the problem of wet-vapour at the turbine.

Since, at ambient temperature (the heat sink), water change phase at lower-than-atmospheric pressure (e.g. 5 kPa at 33 ºC) the condenser must operate under vacuum and a so-called 'deaerator' is needed to remove non-condensable gasses from the feed water or infiltration; moreover, removal of oxygen and carbon dioxide in feed water is always desirable to avoid corrosions in the circuit. Gas solubility in a liquid decreases near the pure-liquid vapour saturation curve, thus deaeration may be achieved by heating the liquid at constant pressure (e.g. adding some vapour) or by making vacuum at constant temperature (e.g. with a small jet of vapour by Venturi suction).

Maximum temperature in a steam power plant is limited by metallurgical constraints to less than 900 K (some 600 ºC), and the maximum pressure depends on the variations to the simple Rankine cycle used, with typical values of 10 MPa (supercritical Rankine cycles surpass 22 MPa). For a simple Rankine cycle with an ideal turbine exiting just at the vapour saturation point, the energy efficiency, with the perfect substance model, is:

                                                                                                        (5.8)

where cp=2 kJ/(kg×K) is an average isobaric thermal capacity of water vapour, and hlv1=2400 kJ/kg the enthalpy of phase change at T1 (Fig. 5.9).

The main variants of the simple Rankine cycle are reheating (a multistage expansion) and regeneration (bleeding some vapour from the middle of the turbine, and before reheating if used, to heat the feed water). These feed-water heaters may be of the open or closed type (Fig. 5.10). In an open feed-water heater, steam extracted at some turbine stage is added to the main feed water stream (that must be previously pressurised to avoid boiling). In a closed feed-water heater, the extracted steam goes through the shell of a shell-and-tubes heat-exchanger and discharges in a lower-pressure heater or the condenser. The mass fraction of vapour to be extracted is designed to be able to heat the main feed-water stream until the saturation temperature of the extracted steam. 


Fig. 5.10. Open and closed feed-water heaters.

Steam turbines (notice that the term is used indistinguishable for the roto-dynamic device and for the whole power plant) are the largest thermal power plants, typically limited to 1000 MW per unit in nuclear power stations, with typical efficiencies from 30% to 40%, although supercritical power plants reach 45% (based on LHV). An advantage of steam turbines, extensive to all external combustion engines, is that any kind of fuel or other heat source may be used, contrary to internal combustion engines, where only fluid fuels, either residual to petroleum distillation but most of the times distillate fluids, can be used. The isentropic efficiency of the turbine is typically 85%, and the electromechanical efficiency of the alternator 98%. The energy and exergy balances in a typical steam power plant are presented in Table 1.

Table 1. Energy and exergy balances in a typical steam power plant
Component
Energy output
Exergy use
Combustion chamber
0
0.30
Boiler tubes heat transfer
0
0.30
Exhaust gasses (chimney)
0.15
0.01
Turbine
0
0.05
Condenser
0
0.03
Water cooling (condenser)
0.55
0.01
Shaft
0.30
0.30

1
1


Kalina cycle
Most heat input/output to/from a plant's working fluid is from variable temperature heat sources/sinks, as the hot combustion gasses and the cold cooling water streams in the normal Rankine cycle. If, instead of using a pure fluid, a mixture were used in a Rankine cycle, due to its variable boiling/condensing temperature, the phase-change heating/cooling could better match the temperature rise/fall in the heating/cooling streams.

The only successful vapour-mixture cycle developed has been the Kalina cycle, proposed by Alexander Kalina in 1984 (the first Kalina power plant, of 3 MW, opened in 1991). Its characteristics are:
·         An ammonia-water mixture (70%-NH3 and 30%-H2O) is used (a large know-how from absorption-refrigeration cycles existed).
·         The turbine exit goes through a distillation and the heavy fraction (40%-NH3 and 60%-H2O) is condensed, and, once pressurised, the two fractions mix before entering the boiler again.
·         Because temperature jumps across heat exchangers can be more uniform, the efficiency increases some 10% for a normal power station, but for special low-temperature applications more than 30%.
·         Because ammonia lowers the boiling point, the Kalina cycle is better suited to low-temperature applications than the Rankine cycle, as in bottoming cycles (see below), geothermal power plants, and so on.

 

 

Combined power cycles

In a Rankine cycle, one single substance, like water, cannot easily match the high-temperature side (e.g. at the temperature of the combustion gasses, 1500 K to 2000 K, it is very difficult to transfer heat to water vapour), and the low-temperature side (to condense water vapour at ambient temperature is difficult because of the very low pressures and densities). The use of two Rankine cycles with different substances has been tried without success (an experimental plant was built with mercury for the top cycle and water for the bottom cycle).

The combination that has reached considerable success is the Brayton-Rankine combined cycle, where the exhaust gasses from a gas turbine are used to supply the heat in the boiler of a vapour turbine operating at not too-high temperatures. The Brayton-Kalina combination may be particularly successful in this respect. Natural-gas-fuelled combined power stations are the rule nowadays because of their low installation cost (some 450 $/kW against 1100 $/kW for coal stations), short-time operations start-up (2 years vs. 3.5 years for coal), and lower environmental impact (nuclear, coal and hydroelectric stations are on hold in Europe and USA), although wider fluctuation in gas price make the choice risky.

The turbodiesel engine can also be considered a combined Diesel-Brayton cycle (Fig.5.11).

Fig. 5.11. Combined power cycles: a) Brayton-Rankine, b) Diesel-Brayton.

There are other combined cycles, still in the developing stage, that show great promise from both the energetics and the environmental aspects, like Graz cycle in Fig. 5.12 (proposed in 1985 by Prof. H. Jericha from Graz University, Austria), where pure oxygen is used as oxidiser instead of air (to avoid NOx-emissions and to have a pure exhaust gas), which on combustion with a fuel (e.g. natural gas) yields only water vapour and carbon dioxide as products, easily separated in a water condenser, ready to CO2 capture, a major concern in the fight against global warming by anthropogenic greenhouse gases. Of course, CO2 capture can be done with all traditional power cycles, e.g. by selective chemical absorption from the exhaust, but it is not competitive with oxy-fuel cycles like Graz's. Application of these air-independent propulsion concepts to submarines is evident.

Fig. 5.12.       The Graz cycle, a new combined gas-steam power cycle burning fuel with pure oxygen, and capturing carbon dioxide.

Propulsion

To propel or impel is to force a body to move forwards, and it requires a push (a force) and an energy expenditure (motive power) to overcome the drag imposed in practice by the nearby objects (if there were no interactions with the surroundings, a body could move without propulsion, as in outer space). Most propulsive systems aim at moving a body over land or through a fluid at constant speed, but all propulsive systems need to accelerate the body at some stage. The power required to propel a body of mass m (e.g. a car) may be expressed as:

                                                                                                   (5.9)

with:

                              (5.10)

where a is the acceleration applied, v the actual speed, q the slope of climbing, cR a tyre-rolling coefficient (typically 10-2), cD a fluid-drag coefficient (typically of order 1, but may drop to 0.1 for streamlined bodies), AF the frontal area projected by the body in the direction of motion, and r the density of the fluid medium.

The idealisation of a long rope being hauled in absence of any other interaction may be a good propulsion paradigm; because of force reciprocity, propulsion may be thought both as a pushing the environment (the rope, in the example) to the rear, and as pulling (the rope) from the front.

As a first model, for a constant-mass vehicle, the momentum balance in the direction of motion may be written as m being the mass of the body,  its acceleration, Tthe thrust force (pull or push on the rope example) and D the drag force resisting the motion (friction with the rope, frictional and pressure drag in a fluid, etc.). For these vehicles, the drag can be measured by towing at constant speed, and the thrust by breaking. For steady motion at speed v, the energy dissipated just by the motion is Dv=Tv, and a propulsion efficiency may be defined to compare this energy with the energy supplied by the engine that cause the motion,  (because it is usually through a shaft), in the way:

                    (shaft-propeller efficiency)                                                                               (5.11)

Notice that all engines consume more energy that what they supply: electrical engines, elastic engines, etc., but particularly thermal engines, where this internal energy conversion efficiency, he, is of the order of 30% to 50% (in electric motors it may be typically 95%).

Most propulsion systems may be considered in this way just power plants that apply a torque, M, to a shaft at a certain rotation speed, w, producing a power =Mw. The shaft then is mechanically coupled to the propeller itself, which may be the friction wheels in land vehicles, or the pressure-reaction blade-propeller in sea and air vehicles.

But there are some propulsion systems based on the change of momentum of a working fluid, notably rockets, jet-engines and water jets, which have inlets and outlets through which some fluid flow. The momentum balance for an open system is :

                                                                                    (5.12)

where m is the mass of the body being propelled (may be changing),  is the sum of externally applied volumetric forces (gravitational or electromagnetic),  the sum of externally applied surface forces at the wall, and  and pe the exit velocity and pressure through any opening of area Ae with an outward normal vector .

Rocket propulsion
For steady horizontal motion of a body with just one exit (rocket) the momentum balance, in the direction of motion, reduces to (see Eq. 5.20):

                                                                                                                       (5.13)

where p0 is the pressure in the surrounding media (acting on the impermeable walls), that may be zero in outer space, and the propulsion efficiency, hp, is defined as the ratio of thrust power, Tv, to whole power emanating from the body, that is the sum of thrust power, Tv (communicated to the surrounding media) and residual kinetic power of the jet relative to the surrounding media,  (communicated to the jet); i.e.:

                                                             (5.14)

where the simplification of considering pe=p0 is always acceptable (it is true for subsonic exits, and for supersonic exits with 'adapted nozzle', and very approximate even for large pressure departures). Equation 5.14 is shown in Fig. 5.13.
Fig. 5.13. Rocket propulsion efficiency vs. flight speed to exit speed ratio.

Most rockets are heat engines because they transform the heating power of a fuel/oxidiser mixture into mechanical energy (the kinetic energy of the jet), and thus an energy efficiency may be defined by:

                                                                                                                                        (5.15)

where hP is the heating power of the fuel/oxidiser mixture per unit mass of reactants. Notice that high energy-efficiency demands high exit speed, demanding also high flight speed to keep a reasonable propulsion efficiency; i.e., rocket propulsion is inefficient at low speed. Compressed-gas rockets are also heat engines, but new magnetohydrodynamic and electrodynamic rockets are not.

Air-breathing propulsion
For steady horizontal motion of a body with one inlet and one exit (like a jet engine) the momentum balance, in the direction of motion, reduces to (see 5. 13):

                                                                                     (5.16)

where the approximation to neglect the additional mass flow rate of fuel is introduced. The propulsion efficiency, defined as for rockets, becomes:

                                                    (5.17)

where the approximation to consider the inlet speed equal to the flight speed has been introduced. Notice that the thrust, , may be obtained with a large Dv and small flow rate or vice versa, but the propulsion efficiency, hp=2/(2+Dv/v), shows that it is better to have a small Dv and a large , what has been implemented in practice, where turbofans of high bypass have superseded simple turbojets.

Similarly to chemical rockets, air-breathing engines (turbojets) are heat engines that transform the heating power of a fuel into mechanical energy (the kinetic energy of the exiting jet minus the one at the inlet), and thus an energy efficiency may be defined by:

                                                                                                                           (5.18)

where hP is the heating power of the fuel/oxidiser mixture per unit mass of fuel. Notice that the energy efficiency is based on thermal-to-mechanical conversion, and the propulsion efficiency is based on mechanical-to-propulsive ratio, thus, the efficiency using the fuel for propulsion is hehp. The overall energy balance of the air-breathing engine may help to identify every term; in axes fixed to the surroundings:

                                                                                                                  (5.19)

with:
                                            (5.20)

 being the chemical enthalpy released,  the propulsive power,  the kinetic energy of the exit jet, and  the thermal enthalpy of the exit jet.

It is important to compare air-breathing propulsion with rocket propulsion for the same fuel expenditure (notice that for rockets , whereas for turbojets  is of the order of 2% of ). The quotient of respective thrusts is:

                                                                             (5.21)

showing that air-breathing engines are much better than rockets (if there is air!) because typical values are ve,turbojet~2×102 m/s, and verockett~5×103 m/s.

 

Cogeneration

Cogeneration is the term coined for the combined generation of work and heat (and, by extension, also of cold). Any heat engine must reject a lot of heat to the environment, but this energy has no exergy (i.e. utility) in the ideal heat engine. Practical engines however reject useful heat in the exhaust gasses and in the cooling circuit, that could be directly used (as when heating the car cabin with the engine cooling water), or the engine modified to provide more useful heat at a small expense in work output.

Without cogeneration, the work needed was taken from the electrical grid, and the heat needed was generated locally in a boiler (heat is much more difficult to transport than work). Cogeneration is provided by using a heat engine (a steam turbine, a gas turbine or a reciprocating engine) to generate work and, at the same time, heat (usually in a exhaust gasses boiler). The great advantage of cogeneration is the energy saving; an additional advantage may be the autonomy gained by self-production, and the main drawback is that the ratio of work-to-heat generation is rather stiff, none can be easily accumulated, and the actual need of work and heat may vary a lot with time (that was one of the main advantages of central production: the levelling of the averaged demand). To stiffness of a cogeneration plant is relaxed by the possibility to send excess work-power to the electrical grid, a convenience that has been enforce by public authorities on account of the social benefit that saving in primary energy resources (fossil fuels) cogeneration brings.

Efficiencies of cogeneration plants are high, but care must be paid not to mix work and heat values (everybody knows that work is usually two or three times more expensive than heat, i.e. the output should be consistently measured in exergy, so that the, instead of W+Q, the output value is W+Q(1-T0/T), where T is the temperature at which heat is delivered.

Reciprocating engine cogeneration
The useful heat-to-work ratio is around 0.6. It is only used for direct heating with the exhaust gasses or the cooling water, but not to generate vapour.

Gas turbine cogeneration
The useful heat-to-work ratio is around 2. It does not require changes in the power plant, it is a versatile cogeneration system, and its use is expanding in most industries: textile, chemical, food. Because the exhaust gasses are relatively hot (some 700 K), it is suitable for high temperature applications.

Vapour turbine cogeneration
The useful heat-to-work ratio is around 7 (suitable for the iron industry, but still low for the ceramic and glass industry). It directly provides useful vapour, either by extracting some of it at an intermediate stage in the turbine, or by not expanding to the low pressures of ambient temperatures but to higher than atmospheric pressures (it is called a back-pressure turbine), to render the whole turbine outlet vapour useful.

 

Efficiency in power generation with heat engines and other generators

Power cannot be generated; only converted from one form to another; what is implied is the generation of some useful power at the expense of some other less-convenient power. But recall that we have restricted here the term ‘power’ to mechanical or electrical power (i.e. excluding thermal power), so that power generation can be seen as the generation of electricity (or equivalent power) from any other energy source: thermal, chemical, nuclear, etc.

Power generation efficiency can be defined as “useful output power divided by input power”, but it is not rigorous enough since at least two choices exist for the evaluation of input power: a) heat-equivalent power, and b) work-equivalent power. Table 2 presents typical values of power generation efficiencies using the raw input power (choice a), the most commonly used, although the net input power criterion, i.e. the exergy or available energy of the raw energy source, would give a sounder measure of the ‘technological efficiency’ of the power plant.

Table 2. Some power generation efficiencies.
Energy source
Typical efficiency [%]
Typical range [%]
Photovoltaic
10
5..15
Solar thermal
15
10..25
Gas turbine
30
15..38
Spark Ignition ICE
30
25..35
Nuclear
33
32..35
Steam turbine
33
25..39
Wind turbine
40
30..50
Compression Ignition ICE
40
35..49
Fuel cell
45
40..70
Combined GT-ST
50
45..60
Hydroelectrical
85
70..90

It is without saying that power generation efficiency is not the only criterion, neither the most important, to quantify engine excellence. The generic goal of maximum power at minimum cost, should include the cost of design (e.g. new technologies), cost of manufacture (e.g. new materials), cost of implementation (e.g. size and weight), cost of operation (e.g. specific fuel consumption, but also pollutant emissions and noise level), cost of maintenance (reliability), and even the cost for disposal.

Problem 1:
Consider a piston-cylinder device, holding initially 500 cm3 of ambient air, which is to be quickly compressed to 50 cm3. We aim at analysing the four following cyclic processes (find the energy exchanges and the energy efficiency):
a)      After the compression, 1 kJ of heat is added at constant volume, and then a quick expansion follows, ending with a constant volume heat release.
b)      After the compression, 1 kJ of heat is added at constant volume, and then a quick expansion follows, ending with a constant pressure heat release.
c)      After the compression, 1 kJ of heat is added at constant pressure, and then a quick expansion follows, ending with a constant volume heat release.
d)      After the compression, 1 kJ of heat is added at constant pressure, and then a quick expansion follows, ending with a constant pressure heat release.

Solution
a)      After the compression, 1 kJ of heat is added at constant volume, and then a quick expansion follows, ending with a constant volume heat release.
            Let us first work out the initial compression, common to all four cycles.

      From 1 to 2. It is a control-mass problem, since in air-standard cycles the mass of air is assumed to remain all the time within the cylinder. The energy balance is thus DE=W+Q. The enclosed mass, assuming p0=105 Pa and T0=288 K, is =105∙500∙10-6/(287∙288)=0,61 g. After the isentropic compression, the state is V2=50 cm3,p2=p1(V1/V2)g=105∙(500/50)1.4=2.5 MPa, T2=T1(V1/V2)g-1=288∙(500/50)0.4=723 K. The work the gas receives is W12=DE=mcv(T2-T1)=0.61∙10-3∙ (1000-287)(723-288)=190 J, and the heat is null, Q12=0. Notice that W12 is not the work that must be done; ambient pressure contributes, too.

      Let start now with the first cycle described, that corresponds to the well-known Otto cycle, represented in Fig. 1.
Fig. 1. Otto cycle.

      From 2 to 3, Q23=1 kJ is received with V3=V2 (i.e. W23=0), and, from the energy balance DE=W+Q=mcv(T3-T2), T3=T2+Q23/(mcv)=723+1000/(0.61∙10-3∙713)=3020 K, too high for practical combustion engines, that only reach 2000 K, but this is a simple academic exercise. Furthermore, p3=p2T3/T2=2.5∙106∙3020/723=11 MPa.

      From 3 to 4 there is an isentropic expansion, from V3=50 cm3 to V4=500 cm3, so that p4=p3(V3/V4)g=11∙106∙(50/500)1.4=418 kPa, T4=T3(V3/V4)g-1=3020∙(50/500)0.4=1210 K. The work the gas delivers is W34=DE=mcv(T4-T3)=0.61∙10-3∙713(1210-3020)=-790 J, and the heat is null, Q34=0. Notice again that W34 is not the work delivered to the shaft; ambient pressure resists, too.

      From 4 to 1, heat is released at constant volume (i.e. W41=0), and, from the energy balance DE=W+Q=mcv(T1-T4)=0.61∙10-3∙713(288-1210)=-400 J. Table 1 shows the summary of energy exchanges.

Table 1. Summary of energy exchanges.
Otto
DE
W
Q
1-2
190
190
0
2-3
1000
0
1000
3-4
-790
-790
0
4-1
-400
0
-400
cycle
0
-600
600

      The energy efficiency of a power cycle is defined as net work delivered divided by heat input, h=Wnet/Qin, that here takes the value:

                                    

      which can be compared with the well-known efficiency of the ideal Otto cycle: h=1-1/rg-1=1-1/100,4=0,60, being the compression ratio r=V1/V2 (with r=10 in this case).

b)      After the compression, 1 kJ of heat is added at constant volume, and then a quick expansion follows, ending with a constant pressure heat release.
Fig. 2. New cycle.

      From 1 to 2 has been worked at the very beginning: p1=105 Pa, T1=288 K, p2=2,5 MPa, T2=723 K, Q12=and W12=190 J.

      From 2 to 3 is the same as in the Otto cycle: p2=2,5 MPa, T2=723 K, p3=11 MPa, T3=3020 K, Q23=kJ and W23=0.

      From 3 to 4 there is an isentropic expansion, from p3=11 MPa to p4=105 Pa, so that V4=V3(p3/p4)1/g =50∙10-6∙(11/0,1)1/1,4=1400 cm3T4=T3(V3/V4)g-1=3020∙(50/1400)0.4=800 K. The work the gas delivers is W34=DE=mcv(T4-T3)=0.61∙10-3∙713(800-3020)=-970 J, and the heat is null, Q34=0.

      From 4 to 1, heat is released at constant pressure. The energy balance DE=W+Q takes the form Q=DH=mcp(T1-T4)=0.61∙10-3∙1000(288-800)=-310 J, whereasW41=-p1(V1-V4)=90 J. Table 2 shows the summary of energy exchanges.

Table 2. Summary of energy exchanges.

DE
W
Q
1-2
190
190
0
2-3
1000
0
1000
3-4
-970
-970
0
4-1
-220
90
-310
cycle
0
-690
690

      The energy efficiency now takes the value:

                                    

c)      After the compression, 1 kJ of heat is added at constant pressure, and then a quick expansion follows, ending with a constant volume heat release.

            This corresponds to the ideal Diesel cycle, depicted in Fig. 3.
Fig. 3. Diesel cycle.

      From 1 to 2 has been worked at the very beginning: p1=105 Pa, T1=288 K, p2=2,5 MPa, T2=723 K, Q12=and W12=190 J.

      From 2 to 3, Q23=1 kJ is received at constant pressure. The energy balance DE=W+Q takes the form Q=DH=mcp(T3-T2), so thatT3=T2+Q23/(mcp)=723+1000/(0.61∙10-3∙1000)=2370 K and V3=V2T3/T2=50∙10-6∙2370/723=164∙10-6 m3, with a work (output) of W23=-p2(V3-V2)=-290 J.

      From 3 to 4 there is an isentropic expansion, from V3=164 cm3 to V4=500 cm3, so that p4=p3(V3/V4)g=2,5∙106∙(164/500)1.4=525 kPa, T4=T3(V3/V4)g-1=2370∙(164/500)0.4=1520 K. The work the gas delivers is W34=DE=mcv(T4-T3)=0.61∙10-3∙713(1520-2370)=-370 J, and the heat is null, Q34=0.

      From 4 to 1, heat is released at constant volume (i.e. W41=0), and, from the energy balance DE=W+Q=mcv(T1-T4)=0.61∙10-3∙713(288-1520)=-530 J. Table 3 shows the summary of energy exchanges.

Table 3. Summary of energy exchanges.
Diesel
DE
W
Q
1-2
190
190
0
2-3
710
-290
1000
3-4
-370
-370
0
4-1
-530
0
-530
cycle
0
-470
470

      The energy efficiency here takes the value:

                                    

      which can be compared with the well-known efficiency of the ideal Diesel cycle:

                                   

      r being the compression ratio r=V1/V2 as for the Otto cycle (r=10), and rc being the cut-off ratio, rc=V3/V2=3,3.

d)      After the compression, 1 kJ of heat is added at constant pressure, and then a quick expansion follows, ending with a constant pressure heat release.

      This corresponds to the ideal Brayton cycle, depicted in Fig. 4.
Fig. 4. Brayton cycle.

      From 1 to 2 has been worked at the very beginning: p1=105 Pa, T1=288 K, p2=2,5 MPa, T2=723 K, Q12=and W12=190 J.

      From 2 to 3, Q23=1 kJ is received at constant pressure as in the Diesel cycle above, yielding T3=2370 K, V3=164∙10-6 m3, and W23=-290 J.

      From 3 to 4 there is an isentropic expansion, from p3=p2=2,5 MPa to p4=105 Pa, so that V4=V3(p3/p4)1/g =164∙10-6∙(2,5/0,1)1/1,4=1640 cm3T4=T3(V3/V4)g-1=2370∙(164/1640)0.4=940 K. The work the gas delivers is W34=DE=mcv(T4-T3)=0.61∙10-3∙713(940-2370)=-620 J, and the heat is null, Q34=0.

      From 4 to 1, heat is released at constant pressure. The energy balance DE=W+Q takes the form Q=DH=mcp(T1-T4)=0.61∙10-3∙1000(288-940)=-400 J, whereasW41=-p1(V1-V4)=114 J. Table 4 shows the summary of energy exchanges.

Table 4. Summary of energy exchanges.
Brayton
DE
W
Q
1-2
189
189
0
2-3
714
-286
1000
3-4
-619
-619
0
4-1
-284
114
-398
cycle
0
-602
602

      The energy efficiency here takes the value:

                                    

      which can be compared with the well-known efficiency of the ideal Brayton cycle:

                                    

      p12 being the pressure ratio p12=p2/p1.

Problem2:
A four-cylinder four-stroke 50 kW gasoline engine, running at 3000 rpm, has a compression ratio of 9, 500 cm3 of displacement per cylinder. Evaluate:
a)             The conversion efficiency, deducing it from the basic cycle states.
b)             Maximum pressure and temperature in the cycle, and mean effective pressure.


Solution.                                  
Fig. 1. Sketch of the standard Otto cycle and corresponding piston position.

a)      The conversion efficiency, deducing it from the basic cycle states.
The cold-air-standard model takes as working fluid plain air with constant properties (those at the inlet, i.e. cold), what renders the analysis simple. The energy exchanges for the trapped control mass, m, are W12/m=cv(T2-T1), Q12=0, W23=0, Q23/m=cv(T2-T1), W34/m=cv(T4-T3), Q34=0, W41=0, Q41/m=cv(T1-T4), and, with the isentropic relation,TVg-1=constant, in terms of the compression ratio rV1/V2, the energy efficiency becomes:
        

with a numerical values of he=1-1/90.4=0.58. Taking as reference a Carnot cycle with the same extreme temperatures, the exergy efficiency hx would be:
   
where T3 is computed below.
b)             Maximum pressure and temperature in the cycle, and mean effective pressure.
For the isentropic compression T2=T1rg-1=288×90.4=694 K, p2=p1rg=100×91.4=2.2 MPa. From the mechanical power:
        

where =50 kW, Z=4 cylinders, m=p1V1/(RT1)=100×103×0.5×10-3/(287×288)=0.68×10-3 kg is the control mass per cylinder (neglecting the residual mass), and n=(3000/60)/2=25 s-1 is the control-mass processing rate (i.e. the charge is renovated 25 times per second; remember that only half of the cycles count, in a 4-stroke engine). The maximum temperature obtained from that is T3=1060 K, and then the maximum pressure p3=p2(T3/T2)=2.2×(1060/694)=3.3 MPa. Substitution in the exergy efficiency above, yields hx=0.80.
The mean effective pressure, pme, is the unit work divided by the displacement volume, i.e. the pressure difference that produces the same power: