Vous êtes sur la page 1sur 28

Heat Temperature and thermal expansion

11-29-99 Sections 13.1 - 13.6 We'll shift gears in the course now, moving from the physics of mechanical systems to thermal physics. Temperature scales In the USA, the Fahrenheit temperature scale is used. Most of the rest of the world uses Celsius, and in science it is often most convenient to use the Kelvin scale. The Celsius scale is based on the temperatures at which water freezes and boils. 0C is the freezing point of water, and 100 C is the boiling point. Room temperature is about 20 C, a hot summer day might be 40 C, and a cold winter day would be around -20 C. To convert between Fahrenheit and Celsius, use these equations:

The two scales agree when the temperature is -40. A change by 1.0 C is a change by 1.8 F. The Kelvin scale has the same increments as the Celsius scale (100 degrees between the freezing and boiling points of water), but the zero is in a different place. The two scales are simply offset by 273.15 degrees. The zero of the Kelvin scale is absolute zero, which is the lowest possible temperature that a substance can be cooled to. Several physics formulas involving temperature only make sense when an absolute temperature (a temperature measured in Kelvin) is used, so the fact that the Kelvin scale is an absolute scale makes it very convenient to apply to scientific work. Measuring temperature A device used to measure temperature is called a thermometer, and all thermometers exploit the fact that properties of a material depend on temperature. The pressure in a sealed bulb depends on temperature; the volume occupied by a liquid depends on

temperature; the voltage generated across a junction of two different metals depends on temperature, and all these effects can be used in thermometers. Linear thermal expansion The length of an object is one of the more obvious things that depends on temperature. When something is heated or cooled, its length changes by an amount proportional to the original length and the change in temperature:

The coefficient of linear expansion depends only on the material an object is made from. If an object is heated or cooled and it is not free to expand or contract (it's tied down at both ends, in other words), the thermal stresses can be large enough to damage the object, or to damage whatever the object is constrained by. This is why bridges have expansion joints in them (check this out where the BU bridge meets Comm. Ave.). Even sidewalks are built accounting for thermal expansion. Holes expand and contract the same way as the material around them. Example This is similar to problem 12.20 in the text. Consider a 2 m long brass rod and a 1 m long aluminum rod. When the temperature is 22 C, there is a gap of 1.0 x 10-3 m separating their ends. No expansion is possible at the other end of either rod. At what temperature will the two bars touch?

The change in temperature is the same for both, the original length is known for both, and the coefficients of linear expansion can be found from Table 12.2 in the textbook.

Both rods will expand when heated. They will touch when the sum of the two length changes equals the initial width of the gap. Therefore:

So, the temperature change is:

If the original temperature was 22 C, the final temperature is 38.4 C. Thermal expansion : expanding holes Consider a donut, a flat, two-dimensional donut, just to make things a little easier. The donut has a hole, with radius r, and an outer radius R. It has a width w which is simply w = R - r. What happens when the donut is heated? It expands, but what happens to the hole? Does it get larger or smaller? If you apply the thermal expansion equation to all three lengths in this problem, do you get consistent results? The three lengths would change as follows:

The final width should also be equal to the difference between the outer and inner radii. This gives:

This is exactly what we got by applying the linear thermal expansion equation to the width of the donut above. So, with something like a donut, an increase in temperature causes the width to increase, the outer radius to increase, and the inner radius to increase, with all dimensions obeying linear thermal expansion. The hole expands just as if it's made as the same material as the hole. Volume thermal expansion When something changes temperature, it shrinks or expands in all three dimensions. In some cases (bridges and sidewalks, for example), it is just a change in one dimension that really matters. In other cases, such as for a mercury or alcohol-filled thermometer, it is the change in volume that is important. With fluid-filled containers, in general, it's how the volume of the fluid changes that's important. Often you can neglect any expansion or contraction of the container itself, because liquids generally have a substantially larger coefficient of thermal expansion than do solids. It's always a good idea to check in a given situation, however, comparing the two coefficients of thermal expansion for the liquid and solid involved.

The equation relating the volume change to a change in temperature has the same form as the linear expansion equation, and is given by:

The volume expansion coefficient is three times larger than the linear expansion coefficient.

Heat
12-3-99 Sections 14.1 - 14.6 Temperature, internal energy, and heat The temperature of an object is a measure of the energy per molecule of an object. To raise the temperature, energy must be added; to lower the temperature, energy has to be removed. This thermal energy is internal, in the sense that it is associated with the motion of the atoms and molecules making up the object. When objects of different temperatures are brought together, the temperatures will tend to equalize. Energy is transferred from hotter objects to cooler objects; this transferred energy is known as heat. Specific heat capacity When objects of different temperature are brought together, and heat is transferred from the higher-temperature objects to the lower-temperature objects, the total internal energy is conserved. Applying conservation of energy means that the total heat transferred from the hotter objects must equal the total heat transferred to the cooler objects. If the temperature of an object changes, the heat (Q) added or removed can be found using the equation:

where m is the mass, and c is the specific heat capacity, a measure of the heat required to change the temperature of a particular mass by a particular temperature. The SI unit for specific heat is J / (kg C).

This applies to liquids and solids. Generally, the specific heat capacities for solids are a few hundred J / (kg C), and for liquids they're a few thousand J / (kg C). For gases, the same equation applies, but there are two different specific heat values. The specific heat capacity of a gas depends on whether the pressure or the volume of the gas is kept constant; there is a specific heat capacity for constant pressure, and a specific heat capacity for constant volume. Example 0.300 kg of coffee, at a temperature of 95 C, is poured into a room-temperature steel mug, of mass 0.125 kg. Assuming no energy is lost to the surroundings, what does the temperature of the mug filled with coffee come to? Applying conservation of energy, the total change in energy of the system must be zero. So, we can just add up the individual energy changes (the Q's) and set the sum equal to zero. The subscript c refers to the coffee, and m to the mug.

Note that room temperature in Celsius is about 20. Re-arranging the equation to solve for the final temperature gives:

The temperature of the coffee doesn't drop by much because the specific heat of water (or coffee) is so much larger than that of steel. This is too hot to drink, but if you leave it heat will be transferred to the surroundings and the coffee will cool. Changing phase; latent heat Funny things happen when a substance changes phase. Heat can be transferred in or out without any change in temperature, because of the energy required to change phase. What is happening is that the internal energy of the substance is changing, because the relationship between neighboring atoms and molecules changes. Going from solid to liquid, for example, the solid phase of the material might have a

particular crystal structure, and the internal energy depends on the structure. In the liquid phase, there is no crystal structure, so the internal energy is quite different (higher, generally) from what it is in the solid phase. The change in internal energy associated with a change in phase is known as the latent heat. For a liquid-solid phase change, it's called the latent heat of fusion. For the gasliquid phase change, it's the latent heat of vaporization, which is generally larger than the latent heat of fusion. Latent heats are relatively large compared to the heat required to change the temperature of a substance by 1 C.

If you use the sum-of-all-the-Q's equals zero equation, you have to be careful with the heat associated with something changing phase because you need to put it in with the appropriate sign. If heat is going into a substance changing phase, such as when it's melting or boiling, the Q is positive; if heat is being removed, such as when it's freezing or condensing, the Q is negative. We don't have to worry about the signs for the heat required to change temperature, because the sign is already built in to the change in temperature. The textbook use an alternate approach (see example 14-5 on page 422). In that method, all the heat losses are set equal to all the heat gains, with everything (even the changes in temperature) going in with positive signs. Note that a change in phase takes place only under the right conditions. Water, for example, doesn't freeze at 10 C, at least not at atmospheric pressure. If you had water at that temperature, you would first need to cool it to the melting point, 0 C, before it would start to freeze. If you're putting in heat from an outside source, the sum-of-all-the-Q's equation becomes:

Heat transfer
12-6-99 Sections 14.7-14.9 Heat transfer

There are three basic ways in which heat is transferred. In fluids, heat is often transferred by convection, in which the motion of the fluid itself carries heat from one place to another. Another way to transfer heat is by conduction, which does not involve any motion of a substance, but rather is a transfer of energy within a substance (or between substances in contact). The third way to transfer energy is by radiation, which involves absorbing or giving off electromagnetic waves. Convection Heat transfer in fluids generally takes place via convection. Convection currents are set up in the fluid because the hotter part of the fluid is not as dense as the cooler part, so there is an upward buoyant force on the hotter fluid, making it rise while the cooler, denser, fluid sinks. Birds and gliders make use of upward convection currents to rise, and we also rely on convection to remove ground-level pollution. Forced convection, where the fluid does not flow of its own accord but is pushed, is often used for heating (e.g., forced-air furnaces) or cooling (e.g., fans, automobile cooling systems). Conduction When heat is transferred via conduction, the substance itself does not flow; rather, heat is transferred internally, by vibrations of atoms and molecules. Electrons can also carry heat, which is the reason metals are generally very good conductors of heat. Metals have many free electrons, which move around randomly; these can transfer heat from one part of the metal to another. The equation governing heat conduction along something of length (or thickness) L and cross-sectional area A, in a time t is:

k is the thermal conductivity, a constant depending only on the material, and having units of J / (s m C). Copper, a good thermal conductor, which is why some pots and pans have copper bases, has a thermal conductivity of 390 J / (s m C). Styrofoam, on the other hand, a good insulator, has a thermal conductivity of 0.01 J / (s m C). Example

Consider what happens when a layer of ice builds up in a freezer. When this happens, the freezer is much less efficient at keeping food frozen. Under normal operation, a freezer keeps food frozen by transferring heat through the aluminum walls of the freezer. The inside of the freezer is kept at -10 C; this temperature is maintained by having the other side of the aluminum at a temperature of -25 C. The aluminum is 1.5 mm thick. Let's take the thermal conductivity of aluminum to be 240 J / (s m C). With a temperature difference of 15, the amount of heat conducted through the aluminum per second per square meter can be calculated from the conductivity equation:

This is quite a large heat-transfer rate. What happens if 5 mm of ice builds up inside the freezer, however? Now the heat must be transferred from the freezer, at -10 C, through 5 mm of ice, then through 1.5 mm of aluminum, to the outside of the aluminum at -25 C. The rate of heat transfer must be the same through the ice and the aluminum; this allows the temperature at the ice-aluminum interface to be calculated. Setting the heat-transfer rates equal gives:

The thermal conductivity of ice is 2.2 J / (s m C). Solving for T gives:

Now, instead of heat being transferred through the aluminum with a temperature difference of 15, the difference is only 0.041. This gives a heat transfer rate of:

With a layer of ice covering the walls, the rate of heat transfer is reduced by a factor of more than 300! It's no wonder the freezer has to work much harder to keep the food cold. Radiation The third way to transfer heat, in addition to convection and conduction, is by radiation, in which energy is transferred in the form of electromagnetic waves. We'll talk about electromagnetic waves in a lot more detail in PY106; an electromagnetic wave is basically an oscillating electric and magnetic field traveling through space at

the speed of light. Don't worry if that definition goes over your head, because you're already familiar with many kinds of electromagnetic waves, such as radio waves, microwaves, the light we see, X-rays, and ultraviolet rays. The only difference between the different kinds is the frequency and wavelength of the wave. Note that the radiation we're talking about here, in regard to heat transfer, is not the same thing as the dangerous radiation associated with nuclear bombs, etc. That radiation comes in the form of very high energy electromagnetic waves, as well as nuclear particles. The radiation associated with heat transfer is entirely electromagnetic waves, with a relatively low (and therefore relatively safe) energy. Everything around us takes in energy from radiation, and gives it off in the form of radiation. When everything is at the same temperature, the amount of energy received is equal to the amount given off. Because there is no net change in energy, no temperature changes occur. When things are at different temperatures, however, the hotter objects give off more energy in the form of radiation than they take in; the reverse is true for the colder objects. The amount of energy an object radiates depends strongly on temperature. For an object with a temperature T (in Kelvin) and a surface area A, the energy radiated in a time t is given by the Stefan-Boltzmann law of radiation:

The constant e is known as the emissivity, and it's a measure of the fraction of incident radiation energy is absorbed and radiated by the object. This depends to a large extent on how shiny it is. If an object reflects a lot of energy, it will absorb (and radiate) very little; if it reflects very little energy, it will absorb and radiate quite efficiently. Black objects, for example, generally absorb radiation very well, and would have emissivities close to 1. This is the largest possible value for the emissivity, and an object with e = 1 is called a perfect blackbody, Note that the emissivity of an object depends on the wavelength of radiation. A shiny object may reflect a great deal of visible light, but it may be a good absorber(and therefore emitter) of radiation of a different wavelength, such as ultraviolet or infrared light. Note that the emissivity of an object is a measure of not just how well it absorbs radiation, but also of how well it radiates the energy. This means a black object that absorbs most of the radiation it is exposed to will also radiate energy away at a higher rate than a shiny object with a low emissivity.

The Stefan-Boltzmann law tells you how much energy is radiated from an object at temperature T. It can also be used to calculate how much energy is absorbed by an object in an environment where everything around it is at a particular temperature :

The net energy change is simply the difference between the radiated energy and the absorbed energy. This can be expressed as a power by dividing the energy by the time. The net power output of an object of temperature T is thus:

Heat transfer in general We've looked at the three types of heat transfer. Conduction and convection rely on temperature differences; radiation does, too, but with radiation the absolute temperature is important. In some cases one method of heat transfer may dominate over the other two, but often heat transfer occurs via two, or even all three, processes simultaneously. A stove and oven are perfect examples of the different kinds of heat transfer. If you boil water in a pot on the stove, heat is conducted from the hot burner through the base of the pot to the water. Heat can also be conducted along the handle of the pot, which is why you need to be careful picking the pot up, and why most pots don't have metal handles. In the water in the pot, convection currents are set up, helping to heat the water uniformly. If you cook something in the oven, on the other hand, heat is transferred from the glowing elements in the oven to the food via radiation.

The gas laws


12-1-99 Sections 13.7 - 13.10 A bit of chemistry Let's delve into some concepts that you might associate with chemistry, but which are equally relevant to physics. The first is the idea of the mole, and Avogadro's number. A mole is a number like a dozen, just a lot bigger. A dozen means 12; a mole means 6.022 x 1023. That number, 6.022 x 1023, is known as Avogadro's number.

You can have a mole of eggs, just as you can have a dozen eggs, but a mole is more useful when it's applied to elements. A mole of aluminum and a mole of lead both have 6.022 x 1023 atoms. The mole of lead has more mass, though, because a mole of an element has a mass in grams equal to the atomic mass listed in the periodic table of elements. A mole of lead, then, has a mass of 207.2 g while a mole of aluminum has a mass of 26.9815 g. A mole of an element has a mass conveniently measured in grams. A single atom, on the other hand, has a mass which is just a small fraction of a gram. A single atom of aluminum, for example, has a mass of:

Note that this is the mass of an average atom of aluminum, accounting for different isotopes. The mass of individual atoms are more conveniently measured in atomic mass units (u). 1 u = 1.66 x 10-24 g = 1.66 x 10-27 kg. For aluminum, then, which has an atomic mass of 26.9815, one mole has a mass of 26.9815 g and an average atom has a mass of 26.9815 u. The gas laws Although we're now familiar with the ideal gas law, which relates the pressure, volume, and temperature of an ideal gas in one compact equation, it's useful to spend a few minutes on the history of gas laws. Three names in particular are associated with gas laws, those being Robert Boyle (1627 - 1691), Jacques Charles (1746-1823), and J.L. Gay-Lussac (1778-1850). Boyle showed that for a fixed amount of gas at constant temperature, the pressure and volume are inversely proportional to one another. In other words: Boyle's law : PV = constant. In Charles' law, it is the pressure that is kept constant. Under this constraint, the volume is proportional to the temperature. This can be expressed as : Charles' law : V1 / T1 = V2 / T2 When the volume is kept constant, it is the pressure of the gas that is proportional to temperature:

Gay-Lussac's law : P1 / T1 = P2 / T2 All of the above laws are combined in the ideal gas law. Before stating that, we should summarize what constitutes an ideal gas. An ideal gas An ideal gas has a number of properties; real gases often exhibit behavior very close to ideal. The properties of an ideal gas are: 1. An ideal gas consists of a large number of identical molecules. 2. The volume occupied by the molecules themselves is negligible compared to the volume occupied by the gas. 3. The molecules obey Newton's laws of motion, and they move in random motion. 4. The molecules experience forces only during collisions; any collisions are completely elastic, and take a negligible amount of time. The ideal gas law An ideal gas is an idealized model of real gases; real gases follow ideal gas behavior if their density is low enough that the gas molecules don't interact much, and when they do interact they undergo elastic collisions, with no loss of kinetic energy. The behavior of an ideal gas, that is, the relationship of pressure (P), volume (V), and temperature (T), can be summarized in the ideal gas law: Ideal gas law : PV = nRT where n is the number of moles of gas, and R = 8.31 J / (mol K) is known as the universal gas constant. An alternate way to express the ideal gas law is in terms of N, the number of molecules, rather than n, the number of moles. N is simply n multiplied by Avogadro's number, so the ideal gas law can be written as:

We'll come back to the ideal gas law, but let's back up a little and get a feel for pressure on the molecular level. This gets us into the theory of the motion of gas molecules, known as the kinetic theory of gases.

Kinetic theory
12-1-99 Sections 13.11 - 13.15 We're now going to draw on much of what we've learned in this course to understand the motion of gases at the molecular level. We'll examine the ideal gas law from the perspective of physics, and we'll come to a deeper understanding of what temperature is. Kinetic theory Consider a cube-shaped box, each side of length L, filled with molecules of an ideal gas. A molecule of ideal gas is like a bouncy rubber ball; whenever it's involved in a collision with a wall of the box, it rebounds with the same kinetic energy it had before hitting the wall. Similarly, if ideal gas molecules collide, the collisions are elastic, so no kinetic energy is lost. Now consider one of these ideal gas molecules in this box, with a mass m and velocity v. If this molecule bounces off one of the walls perpendicular to the x-direction, the y and z components of the molecule's velocity are unaffected, and the x-component of velocity reverses. The molecule maintains the same speed, because the collsion is elastic. How much force, on average, does it exert on the wall? To answer this we just have to think about momentum, and impulse. Momentum is a vector, so if the particle reverses its x-component of momentum there is a net change in momentum of 2 m vx. The magnitude of the average force exerted by the wall to produce this change in momentum can be found from the impulse equation:

If we're dealing with the average force, the time interval is just the time between collisions, which is the time it takes for the molecule to make a round trip in the box and come back to hit the same wall. This time simply depends on the distance travelled in the x-direction (2L), and the x-component of velocity; the time between collisions with that wall is 2L / vx. Solving for the average force gives:

This is the magnitude of the average force exerted by the wall on the molecule, as well as the magnitude of the average force exerted by the molecule on the wall; the forces are equal and opposite. This is just the average force exerted on one wall by one molecule selected at random. To find the total force, we need to include all the other molecules, which travel at a wide range of speeds. The distribution of speeds follows a curve that looks a lot like a Bell curve, with the peak of the distribution increasing as the temperature increases. The shape of the speed distribution is known as a Maxwellian distribution curve, named after James Clerk Maxwell, the Scottish physicist who first worked it out. To find the total force on the wall, then, we can just add up all the individual forces:

Multiplying and dividing the right-hand side by N, the number of molecules in the box, gives:

The term in square brackets is simply an average...we're adding up the square of the xcomponent of velocity for all the molecules and dividing by the number of molecules. This average is known as the root-mean-square average, symbolized by rms.

Therefore:

Consider now how this average x velocity compares to the average velocity. For any molecule, the velocity can be found from its components using the Pythagorean theorem in three dimensions:

For a randomly-chosen molecule, the x, y, and z components of velocity may be similar but they don't have to be. If we take an average over all the molecules in the box, though, then the average x, y, and z speeds should be equal, because there's no reason for one direction to be preferred over another. So:

The force exerted by the molecules on a wall of the box can then be expressed in terms of the average velocity, rather than the average x component of velocity.

The pressure exerted by the gas on each wall is simply the force divided by the area of a wall. Therefore:

Rearranging things a little gives:

This equation has many of the same variables as the ideal gas law, PV = NkT. This means that:

This is a very important result, because it tells us something fundamental about temperature. The absolute temperature of an ideal gas is proportional to the average kinetic energy per gas molecule. If changes in pressure and/or volume result in changes in temperature, it means the average kinetic energy of the molecules has been changed. The internal energy of an ideal gas The result above says that the average translational kinetic energy of a molecule in an ideal gas is 3/2 kT. For a gas made up of single atoms (the gas is monatomic, in other words), the translational kinetic energy is also the total internal energy. Rotational kinetic energy can be ignored, because the atoms are so small that the moment of inertia is negligible.There is also no energy associated with bonds between atoms in molecules, because there are no bonds in a monatomic gas. For a monatomic ideal gas, then, the total internal energy (U) is simply:

Note that gases made up of molecules with more than one atom per molecule have more internal energy than 3/2 nRT, because energy is associated with the bonds and the vibration of the molecules. Another interesting result involves the equipartition of energy, which says that each contributer to the internal energy contributes an equal

amount of energy. For a monatomic ideal gas, each of the three directions (x, y, and z) contribute 1/2 kT per molecule, for a total of 3/2 kT per molecule. For a diatomic molecule, the three translation direction contribute 1/2 kT per molecule, and for each molecule there are also two axes of rotation, contributing rotational kinetic energy in the amount of 1/2 kT each. This amounts to a total internal energy of U = 5/2 NkT for a diatomic gas, and a polyatomic gas has contributions from three translation direction and three axes of rotation, giving U = 6/2 NkT, or 3NkT. Diffusion Physics is full of cases where similar behavior in different kinds of systems can be described by similar equations. Diffusion, for example, the flow of a substance from a region of higher concentration to a region of lower concentration, is a similar process to thermal conductivity, which involves heat flowing from a higher temperature region to a lower temperature region. The similarities can be seen in the equations:

Both equations involve a length, a time, a cross-sectional area, and a constant. D is the diffusion constant, while k is the thermal conductivity; both depend on what the flow is passing through. In addition, both equations involve differences; temperature difference for heat flow, and a concentration difference in the case of the mass flow. Always look for parallels. If you can relate one system back to another system you already understand, you'll be able to figure out the new system much more easily.

The first law of thermodynamics


12-8-99 Sections 15.1 - 15.4 Thermodynamics Thermodynamics is the study of systems involving energy in the form of heat and work. A good example of a thermodynamic system is gas confined by a piston in a cylinder. If the gas is heated, it will expand, doing work on the piston; this is one example of how a thermodynamic system can do work.

Thermal equilibrium is an important concept in thermodynamics. When two systems are in thermal equilibrium, there is no net heat transfer between them. This occurs when the systems are at the same temperature. In other words, systems at the same temperature will be in thermal equilibrium with each other. The first law of thermodynamics relates changes in internal energy to heat added to a system and the work done by a system. The first law is simply a conservation of energy equation:

The internal energy has the symbol U. Q is positive if heat is added to the system, and negative if heat is removed; W is positive if work is done by the system, and negative if work is done on the system. We've talked about how heat can be transferred, so you probably have a good idea about what Q means in the first law. What does it mean for the system to do work? Work is simply a force multiplied by the distance moved in the direction of the force. A good example of a thermodynamic system that can do work is the gas confined by a piston in a cylinder, as shown in the diagram.

If the gas is heated, it will expand and push the piston up, thereby doing work on the piston. If the piston is pushed down, on the other hand, the piston does work on the gas and the gas does negative work on the piston. This is an example of how work is done by a thermodynamic system. An example with numbers might make this clearer. An example of work done

Consider a gas in a cylinder at room temperature (T = 293 K), with a volume of 0.065 m3. The gas is confined by a piston with a weight of 100 N and an area of 0.65 m2. The pressure above the piston is atmospheric pressure. (a) What is the pressure of the gas? This can be determined from a free-body diagram of the piston. The weight of the piston acts down, and the atmosphere exerts a downward force as well, coming from force = pressure x area. These two forces are balanced by the upward force coming from the gas pressure. The piston is in equilibrium, so the forces balance. Therefore:

Solving for the pressure of the gas gives:

The pressure in the gas isn't much bigger than atmospheric pressure, just enough to support the weight of the piston. (b) The gas is heated, expanding it and moving the piston up. If the volume occupied by the gas doubles, how much work has the gas done? An assumption to make here is that the pressure is constant. Once the gas has expanded, the pressure will certainly be the same as before because the same freebody diagram applies. As long as the expansion takes place slowly, it is reasonable to assume that the pressure is constant. If the volume has doubled, then, and the pressure has remained the same, the ideal gas law tells us that the temperature must have doubled too. The work done by the gas can be determined by working out the force applied by the gas and calculating the distance. However, the force applied by the gas is the pressure times the area, so: W=Fs=PAs and the area multiplied by the distance is a volume, specifically the change in volume of the gas. So, at constant pressure, work is just the pressure multiplied by the change in volume:

This is positive because the force and the distance moved are in the same direction, so this is work done by the gas. The pressure-volume graph As has been discussed, a gas enclosed by a piston in a cylinder can do work on the piston, the work being the pressure multiplied by the change in volume. If the volume doesn't change, no work is done. If the pressure stays constant while the volume changes, the work done is easy to calculate. On the other hand, if pressure and volume are both changing it's somewhat harder to calculate the work done. As an aid in calculating the work done, it's a good idea to draw a pressure-volume graph (with pressure on the y axis and volume on the x-axis). If a system moves from one point on the graph to another and a line is drawn to connect the points, the work done is the area underneath this line. We'll go through some different thermodynamic processes and see how this works. Types of thermodynamic processes There are a number of different thermodynamic processes that can change the pressure and/or the volume and/or the temperature of a system. To simplify matters, consider what happens when something is kept constant. The different processes are then categorized as follows : 1. Isobaric - the pressure is kept constant. An example of an isobaric system is a gas, being slowly heated or cooled, confined by a piston in a cylinder. The work done by the system in an isobaric process is simply the pressure multiplied by the change in volume, and the P-V graph looks like:

2. Isochoric - the volume is kept constant. An example of this system is a gas in a box with fixed walls. The work done is zero in an isochoric process, and the PV graph looks like:

3. Isothermal - the temperature is kept constant. A gas confined by a piston in a cylinder is again an example of this, only this time the gas is not heated or cooled, but the piston is slowly moved so that the gas expands or is compressed. The temperature is maintained at a constant value by putting the system in contact with a constant-temperature reservoir (the thermodynamic definition of a reservoir is something large enough that it can transfer heat into or out of a system without changing temperature). If the volume increases while the temperature is constant, the pressure must decrease, and if the volume decreases the pressure must increase. 4. Adiabatic - in an adiabatic process, no heat is added or removed from the system. The isothermal and adiabatic processes should be examined in a little more detail. Isothermal processes In an isothermal process, the temperature stays constant, so the pressure and volume are inversely proportional to one another. The P-V graph for an isothermal process looks like this:

The work done by the system is still the area under the P-V curve, but because this is not a straight line the calculation is a little tricky, and really can only properly be done using calculus.

The internal energy of an ideal gas is proportional to the temperature, so if the temperature is kept fixed the internal energy does not change. The first law, which deals with changes in the internal energy, thus becomes 0 = Q - W, so Q = W. If the system does work, the energy comes from heat flowing into the system from the reservoir; if work is done on the system, heat flows out of the system to the reservoir. Adiabatic processes In an adiabatic process, no heat is added or removed from a system. The first law of thermodynamics is thus reduced to saying that the change in the internal energy of a system undergoing an adiabatic change is equal to -W. Since the internal energy is directly proportional to temperature, the work becomes:

An example of an adiabatic process is a gas expanding so quickly that no heat can be transferred. The expansion does work, and the temperature drops. This is exactly what happens with a carbon dioxide fire extinguisher, with the gas coming out at high pressure and cooling as it expands at atmospheric pressure.

Specific heat capacity of an ideal gas With liquids and solids that are changing temperature, the heat associated with a temperature change is given by the equation:

A similar equation holds for an ideal gas, only instead of writing the equation in terms of the mass of the gas it is written in terms of the number of moles of gas, and use a capital C for the heat capacity, with units of J / (mol K):

For an ideal gas, the heat capacity depends on what kind of thermodynamic process the gas is experiencing. Generally, two different heat capacities are stated for a gas, the heat capacity at constant pressure (Cp) and the heat capacity at constant volume (Cv). The value at constant pressure is larger than the value at constant volume because at constant pressure not all of the heat goes into changing the temperature; some goes into doing work. On the other hand, at constant volume no work is done, so all the heat goes into changing the temperature. In other words, it takes less heat to produce a given temperature change at constant volume than it does at constant pressure, so Cv < Cp. That's a qualitative statement about the two different heat capacities, but it's very easy to examine them quantitatively. The first law says:

We also know that PV = nRT, and at constant pressure the work done is:

Note that this applies for a monatomic ideal gas. For all gases, though, the following is true:

Another important number is the ratio of the two specific heats, represented by the Greek letter gamma (K). For a monatomic ideal gas this ratio is:

Heat engines and the second law

12-10-99 Sections 15.5 - 15.6 The second law of thermodynamics The second law of thermodynamics comes in more than one form, but let's state in a way that makes it obviously true, based on what you've observed from simply being alive. The second law states that heat flows naturally from regions of higher temperature to regions of lower temperature, but that it will not flow naturally the other way. Heat can be made to flow from a colder region to a hotter region, which is exactly what happens in an air conditioner, but heat only does this when it is forced. On the other hand, heat flows from hot to cold spontaneously. Heat engines We'll move on to look at heat engines, which are devices that use heat to do work. A basic heat engine consists of a gas confined by a piston in a cylinder. If the gas is heated, it expands, moving the piston. This wouldn't be a particularly practical engine, though, because once the gas reaches equilibrium the motion would stop. A practical engine goes through cycles; the piston has to move back and forth. Once the gas is heated, moving the piston up, it can be cooled and the piston will move back down. A cycle of heating and cooling will move the piston up and down. A necessary component of a heat engine, then, is that two temperatures are involved. At one stage the system is heated, at another it is cooled. In a full cycle of a heat engine, three things happen: 1. Heat is added. This is at a relatively high temperature, so the heat can be called QH. 2. Some of the energy from that input heat is used to perform work (W). 3. The rest of the heat is removed at a relatively cold temperature (QC).

The following diagram is a representation of a heat engine, showing the energy flow:

An important measure of a heat engine is its efficiency: how much of the input energy ends up doing useful work? The efficiency is calculated as a fraction (although it is often stated as a percentage):

Work is just the input heat minus the rejected heat, so:

Note that this is the maximum possible efficiency for an engine. In reality there will be other losses (to friction, for example) that will reduce the efficiency. Carnot's principle How can an engine achieve its maximum efficiency? It must operate using reversible processes: a reversible process is one in which the system and the surroundings can be returned to state they were in before the process began. If energy is lost to friction during a process, the process is irreversible; if energy is lost as heat flows from a hot region to a cooler region, the process is irreversible. The efficiency of an engine using irreversible processes can not be greater than the efficiency of an engine using reversible processes that is working between the same temperatures. This is known as Carnot's principle, named after Sadi Carnot, a French engineer. For any reversible engine (known as a Carnot engine) operating between two temperatures, TH and TC, the efficiency is given by:

The efficiency is maximized when the cold reservoir is as cold as possible, and the hot temperature is as hot as possible.

The third law The third law of thermodynamics states this : it is impossible to reach absolute zero. This implies that a perpetual motion machine is impossible, because the efficiency will always be less than 1. Refrigerators, air conditioners, etc. A device such as a refrigerator or air conditioner, designed to remove heat from a cold region and transfer it to a hot region, is essentially a heat engine operating in reverse, as the following energy flow diagram shows:

A refrigerator, consisting of a fluid pumped through a closed system, involves a fourstep process. An air conditioner works the same way.
y

Step 1 - The fluid passes through a nozzle and expands into a low-pressure area. Similar to the way carbon dioxide comes out of a fire extinguisher and cools down, the fluid turns into a gas and cools down. This is essentially an adiabatic expansion. Step 2 - The cool gas is in thermal contact with the inner compartment of the fridge; it heats up as heat is transferred to it from the fridge. This takes place at constant pressure, so it's an isobaric expansion. Step 3 - The gas is transferred to a compressor, which does most of the work in this process. The gas is compressed adiabatically, heating it and turning it back to a liquid. Step 4 - The hot liquid passes through coils on the outside of the fridge, and heat is transferred to the room. This is an isobaric compression process.

A refrigerator is rated by something known as the coefficient of performance, which is the ratio of the heat removed from the fridge to the work required to remove it:

The P-V graph for a refrigerator cycle The P-V (pressure-volume) graph is very useful for calculating the work done. For any kind of heat engine or refrigerator (reverse heat engine), the processes involved form a cycle on the P-V graph. The work is the area of the enclosed region on the graph. The diagram for a refrigerator is a little more complicated than this because of the two phase changes involved, but this is basically what it looks like:

Entropy and the second law


12-12-99 Sections 15.7 - 15.12 The second law revisited The second law of thermodynamics is one of the most fundamental laws of nature, having profound implications. In essence, it says this:

The second law - The level of disorder in the universe is steadily increasing. Systems tend to move from ordered behavior to more random behavior. One implication of the second law is that heat flows spontaneously from a hotter region to a cooler region, but will not flow spontaneously the other way. This applies to anything that flows: it will naturally flow downhill rather than uphill. If you watched a film forwards and backwards, you would almost certainly be able to tell which way was which because of the way things happen. A pendulum will gradually lose energy and come to a stop, but it doesn't pick up energy spontaneously; an ice cube melts to form a puddle, but a puddle never spontaneously transforms itself into an ice cube; a glass falling off a table might shatter when it hits the ground, but the pieces will never spontaneously jump back together to form the glass again. Many processes are irreversible, and any irreversible process increases the level of disorder. One of the most important implications of the second law is that it indicates which way time goes - time naturally flows in a way that increases disorder. The second law also predicts the end of the universe: it implies that the universe will end in a "heat death" in which everything is at the same temperature. This is the ultimate level of disorder; if everything is at the same temperature, no work can be done, and all the energy will end up as the random motion of atoms and molecules. Entropy A measure of the level of disorder of a system is entropy, represented by S. Although it's difficult to measure the total entropy of a system, it's generally fairly easy to measure changes in entropy. For a thermodynamic system involved in a heat transfer of size Q at a temperature T , a change in entropy can be measured by:

The second law of thermodynamics can be stated in terms of entropy. If a reversible process occurs, there is no net change in entropy. In an irreversible process, entropy always increases, so the change in entropy is positive. The total entropy of the universe is continually increasing. There is a strong connection between probability and entropy. This applies to thermodynamic systems like a gas in a box as well as to tossing coins. If you have four pennies, for example, the likelihood that all four will land heads up is relatively small. It is six times more likely that you'll get two heads and two tails. The two heads - two tails state is the most likely, shows the most disorder, and has the highest entropy. Four heads is less likely, has the most order, and the lowest entropy. If you

tossed more coins, it would be even less likely that they'd all land heads up, and even more likely that you'd end up with close to the same number of heads as tails. With a gas in a box, the probability that all the gas molecules are in one corner of the box at the same time is very small (for a typical box full of 1020 molecules or more, incredibly small): this is therefore a low entropy state. It is much more likely that the molecules are randomly distributed around the box, and are moving in random directions; this high disorder state is a considerably higher entropy state. The second law doesn't rule out all the molecules ending up in one corner, but it means it's far more likely that the molecules will be randomly distributed, and to move towards a random distribution from an orderly distribution, as opposed to the other way around.

Vous aimerez peut-être aussi