Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation
The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation
The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation
Ebook2,530 pages21 hours

The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation

Rating: 1 out of 5 stars

1/5

()

Read preview

About this ebook

With over 1,000 pages covering all fundamental and practical HVAC design procedures and methods, this classic reference is packed with details and contains a wealth of information that is of great use to the HVAC designer and practitioner, as well as to the student mastering the intricacies of HVAC fundamentals. Unlike any other handbook of its kind, HVAC provides an in-depth treatment of topics via modular self-contained chapters that serve both as a manual for the experienced professional and as a fundamental reference for others. Each self-contained chapter places emphasis on graphical and tabular presentations of data that are useful for easy understanding of fundamentals and solving problems of design, installation, and operation. You are sure to find everything you need right here in one complete volume!
LanguageEnglish
Release dateMay 1, 2007
ISBN9780831191511
The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation

Related to The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation

Related ebooks

Construction For You

View More

Related articles

Reviews for The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation

Rating: 1 out of 5 stars
1/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Handbook of Heating, Ventilation and Air Conditioning (HVAC) for Design and Implementation - Ali Vedavarz

    FUNDAMENTALS

    This chapter covers the fundamentals of thermodynamics, fluid mechanics, and heat transfer as related to the theory and practice of air conditioning, heating and ventilation. Basic concepts needed for the HVAC professional are presented, while advanced topics are not considered. Excellent resources, such as the ASHRAE Handbook series and many great textbooks in the areas of thermal fluid sciences, treat the details and advanced topics that have been omitted from this focused chapter.

    Fundamentals of Thermodynamics

    Thermodynamics is the study of energy, its transformations, and its relation to states of matter or substance. Thermodynamics deals with equilibrium conditions that are typical of steady state, and any changes are considered to be quasi-equilibrium processes where change occurs slowly and incrementally so as to allow each incremental intermediate state to reach equilibrium before it advances further.

    A pure substance has a homogeneous and invariable chemical composition. It can exist in more than one phase, but chemical composition is the same in all phases. If a substance exists as liquid at saturation temperature and pressure, it is called a saturated liquid. If the temperature of the liquid is lower than the saturation temperature for the existing pressure, it is called either a subcooled liquid (the temperature is lower than the saturation temperature for the given pressure) or a compressed liquid (the pressure is greater than the saturation pressure for the given temperature). When a substance exists as part liquid and part vapor at the saturation temperature, its quality is defined as the ratio of the mass of vapor to the total mass. Quality has meaning only when the substance is in a saturated state; i.e., at saturation pressure and temperature. If a substance exists as vapor at the saturation temperature, it is called saturated vapor. The term dry saturated vapor is used to emphasize that the quality of the substance is 100%. When the vapor is at a temperature greater than the saturation temperature, it is superheated vapor. The pressure and temperature of superheated vapor are independent properties, since the temperature can increase while the pressure remains constant. Gases are highly superheated vapors.

    A property of a substance is any observable characteristic of the material. An intrinsic property of the material is one that does not depend on the shape or volume or mass of the material. Such properties are pressure and temperature, or properties that are usually expressed in the units of per unit mass, such as specific volume. This is in contrast with extrinsic properties which depend on the actual quantity of the material under consideration, such as volume. The thermodynamic state of a substance is defined by listing its intrinsic properties. The most common intrinsic thermodynamic properties are temperature T, pressure p, specific volume v (which is the inverse of density ρ), specific entropy s, specific enthalpy h, and specific internal energy u. It has been established that a thermodynamic state of a substance can be uniquely identified by two independent intrinsic properties. For example, temperature and pressure are such properties, except in the saturation region where the same temperature and pressure can have an infinite number of states corresponding to quality from 0 to 100%. Enthalpy (h) is defined by

    where u = internal energy per unit mass.

    Each property in a given state has only one definite value, and any property always has the same value for a given state, regardless of how the substance arrived at that state. The thermodynamic property entropy s measures the molecular disorder of a system. The more mixed a system, the greater its entropy; conversely, an orderly or unmixed configuration is one of low entropy. Figs. 1-1a and 1-1b schematically show the liquid and vapor states of water using two properties to uniquely identify the states. Fig. 1-1a shows the states in (p, h) coordinates and a line showing the states along a constant temperature is also presented. Fig. 1-1b shows the same information in (T, s) coordinates and a constant pressure line is shown. Within the dome where the saturated states exist it is possible to increase the specific volume (and other intrinsic, per unit mass based properties) even when keeping the pressure and temperature constant (for example line B-A-C in Figs. 1-1a and 1-1b).

    Fig 1-1a. Thermodynamic states of water in p-h coordinates

    Fig 1-1b. Thermodynamic states of water in T-s coordinates

    The specific heats at constant pressure Cp and volume Cv are defined as the heat added to a unit mass at constant pressure or volume, respectively, to cause temperature of the mass to increase by one unit. Mathematically, by using the first law for closed systems and the definition of work (described later in the chapter), these are defined as

    For ideal gases the following holds true rigorously

    and for real gases this is a good engineering approximation. For incompressible substances the two specific heats are the same and the above expressions also holds. Thus, liquids and solids, which are treated as incompressible, are described by Equation (3).

    The basic entity analyzed by thermodynamics is termed a thermodynamic system. A thermodynamic system is an identifiable or specific region in space or an identifiable or specific quantity of matter. It is bounded by the system surface or the system boundaries. The surroundings include everything external to the system, and the system is separated from the surroundings by the system boundaries. These boundaries can be movable or fixed, real or imaginary. There are two types of systems: closed and open. In a closed system, no mass enters or leaves the system. A closed system is thus a fixed mass system. In an open system, mass may enter or leave the system and therefore the boundaries of such a system are imaginary in some parts where the mass is exchanged with the surroundings. Energy can always cross the system boundaries, whether open or closed, except for an isolated closed system, which is one that does not allow energy to cross its boundaries. Almost all HVAC systems that will be studied here fall in the open system category.

    Energy has the capacity for producing an effect on the system and is described differently depending on whether it is stored or is crossing the system boundary. Thermal (internal) energy is the energy possessed by the matter in a system caused by the internal motion of the molecules, the intermolecular forces, and other microscopic mechanisms of energy storage within the material. It is impossible to find the absolute (total) value of internal energy of a material, and it is therefore measured with reference to a standard state in which the value of the internal energy is arbitrarily set to zero. The internal energy is the sum of all these microscopic energies such that when the internal energy of an isolated closed system (fixed mass) changes, the result is a change in the temperature of the system. Internal energy is an intrinsic property of material and its value per unit mass can be used to uniquely define the state of the material. The specific internal energy u is the internal energy per unit mass. Potential energy is the energy possessed by a system due to the elevation of the system:

    Potential energy is not an intrinsic property of material. Kinetic energy is the energy possessed due to the bulk velocity of the flowing material and is expressed as

    where V = velocity of a fluid stream.

    Kinetic energy is not an intrinsic property of material.

    The types of energy that are defined only when they cross the system boundaries are heat and work. Heat (Q) and work (W) cannot be stored and they only exist when energy is being transferred across a system boundary. Heat is the mechanism that transfers energy across the boundary of systems solely through temperature difference. Heat always flows from higher to lower temperatures. Heat is considered positive by convention when energy is added to a system (see Fig. 1-2). Work is the mechanism that transfers energy across the boundary of systems through forces and movement. (Here it has the traditional characterization: it is the product of force and distance.) If the total effect produced in the system can be reduced to the raising of a weight, then nothing but work has crossed the boundary. By convention work is considered positive when energy is removed from a system, i.e., the system does work on its surroundings (see Fig. 1-2). For example, when a system comprising a cylinder and movable piston expands so that the piston is moving outwards, work is done by the system on the surroundings and is therefore considered positive in any analysis of system. Mechanical or shaft work is the work associated with a rotating shaft such as a turbine, air compressor, or internal combustion engine. Flow work is energy associated with the movement of fluid in conjunction with forces of fluid pressure as it crosses the boundary of an open system. It can be more easily understood as the work done by the fluid to push itself against the other fluid particles as it forces itself to enter or exit the system. Work done by entering fluid streams is negative (pushing into the system), while work done by fluid streams exiting the system is positive (pushing the surroundings). The magnitude of flow work per unit mass is determined by the expression

    Fig 1-2. Energy flow in general thermodynamic system

    A property of a system is any observable characteristic of the system. A process is a change in state that can be defined as any change in the properties of a system. A process is described by its initial and final equilibrium states, its path, and the interactions that take place across system boundaries as it goes forth. A cycle is a process or a series of processes wherein the initial and final states of the system are identical. Therefore, at the conclusion of a cycle, all properties have the same value they had at the beginning.

    Conservation of Mass.—Conservation of mass in a closed system is the default since no mass leaves or enters the system. For an open system the conservation of mass indicates that the difference between the mass entering and leaving is equal to the increase of mass in the system. For the general case of multiple flow streams the conservation of mass is written as

    For steady flow processes the above can be modified as

    For flow through a pipe or duct the mass flow rate is related to the velocity by

    It is assumed that the velocity is uniform across the cross-section. If it is not, then V in the above mass flow rate definition is average velocity. A is normal to the direction of the fluid flow.

    First Law of Thermodynamics.—The first law of thermodynamics is often called the law of the conservation of energy. After a system is defined, the conservation of energy states that

    Fig. 1-2 illustrates energy flows into and out of a thermodynamic system. For a closed system this is written as

    For the general case of an open system with multiple mass flows in and out of the system, the energy balance can be written

    The steady flow process is important in engineering applications. Steady flow signifies that no quantities associated with the system vary with time. Consequently,

    Here the addition of the dot on top of the variables m, Q, and W indicates the time derivative (d/dt) which yields the rate of mass flow, the rate of heat flow, and the rate of work done, respectively. In the above two equations the flow work of the entering and exiting streams is indicated by the pv terms, and, therefore, the W in the equations is the work done by moving system boundaries plus shaft work, and any other work not described by these categories. In the above energy conservation ( Equations (12) and (13)) assume that velocity is uniform across the cross section. However, in real fluid systems it is not uniform due to viscosity of the fluid. The change to the kinetic energy term is discussed in the section Fundamentals of Fluid Flow.

    Second Law, Reversibility, and Possible Processes.—The second law of thermodynamics can be expressed in many ways. Here it is being introduced to distinguish and quantify processes that can only proceed in one direction (irreversible) from those that are reversible. It also indicates which processes cannot exist.

    The second law for a closed system is written as

    where the subscripts f and i indicate final and initial state, respectively, and the system entropy S = ms, where m is the mass of the system and s is specific entropy. The temperature T must be in absolute units (Kelvin or Rankine). The equality sign holds for reversible processes and the inequality for irreversible processes. The above form of the second law shows that the entropy of the system either increases or remains same, assuming the heat flow is positive. For adiabatic processes, that is where Q = 0, the right side of either Equation (14) vanishes, clearly indicating that entropy increases if the process is not reversible. For constant temperature processes

    where the equality holds for reversible processes. For an open system the above can be modified as

    It is assumed in Equation (16) that the inlet and outlet properties remain invariant with time. For steady flow systems with invariant inlet and outlet properties the second law can be rewritten in the form

    The open system of Equation (17) can give insights about reversible and irreversible processes if we consider a single input, single output stream process. If the process is adiabatic, then the inlet and outlet entropies will be the same for reversible processes; otherwise the outlet entropy will be greater than the inlet. If the process is isothermal, then the difference between the rate of entropy flowing out and in will be equal to the heat addition rate divided by temperature only if the process is reversible. Adiabatic reversible processes are also termed isentropic because entropy remains same. On a thermodynamic chart where entropy is the horizontal axis, the adiabatic reversible processes will be vertical lines, but the irreversible processes will always veer right (towards increasing entropy) from their starting point. Thus, for any given starting point, the only thermodynamically possible adiabatic processes are those that end in the half plane to the right of the vertical line drawn from the starting point in the thermodynamic chart where specific entropy is on the horizontal axis. Thus, the second law can show processes that are reversible, irreversible and possible, or impossible.

    The second law for cycles is described next. It further refines the concept of possible and impossible series of processes occurring in a cycle.

    Thermodynamic Cycles.—Thermodynamic cycles that make it possible to remove heat from cold spaces and dump the heat in hot ambient spaces are discussed here. These cycles make it possible for air conditioning and refrigeration systems to exist and they form the basis of HVAC engineering.

    The performance of a refrigeration or air conditioning thermodynamic cycle is usually described by a coefficient of performance. COP is defined as the benefit of the cycle (amount of heat removed) divided by the required energy input to operate the cycle, or

    The first law for cycles indicates that

    For the individual processes that make up the cycle, the first law for open systems is used. Since cyclic devices have only one working fluid and the mass flow rate of the fluid is the same through each process due to conservation of mass, the open system Equation (13) can be modified as below for each process in the cycle:

    The kinetic and potential energy terms are usually neglected in comparison to enthalpy in refrigeration and air conditioning thermodynamic cycles because their magnitudes are usually several orders smaller than enthalpy.

    The Carnot cycle is an ideal cycle that is made up of completely reversible processes and operates between two fixed temperatures. This cycle is useful since it is a thermodynamic ideal for refrigeration cycles that need to operate between two temperatures: the temperature of the conditioned space and the temperature of the external (hot) ambient. Also heat is transferred from the cold space to the ambient. The properties of the Carnot cycle are that it can be used as a refrigerator or a heat pump, as well as a work-producing device, depending on the input quantities. Fig. 1-3 shows the Carnot cycle on temperature entropy coordinates. Heat is withdrawn at the constant temperature TR from the region to be refrigerated. Heat is rejected at the constant ambient temperature To. The cycle is completed by two isentropic processes that connect the high and low temperatures at the two extremes of entropy values. From Equation (17) the energy transfers are given by

    Since the cycle is being run to remove the heat QR, the COP of the Carnot cycle will be

    where Wnet work that would be supplied through isentropic processes of the cycle.

    Equation (23) also shows that the COP of the ideal reversible Carnot cycle is a function only of the two absolute temperatures between which it operates.

    Fig 1-3. Carnot cycle

    Fig 1-4. Possible thermodynamic realization of a Carnot cycle using a pure substance

    The second law for cycles states that (1) no refrigerating cycle may have a COP higher than that for a reversible cycle operated between the same temperature limits, and (2) all reversible cycles, when operated between the same temperature limits, have the same COP. Proofs of these statements are not presented here. They can be found in standard thermodynamics textbooks. Although the first law states that net heat flow into the cycle is equal to the net work done by the cycle (see Equation (19)), constraints are placed by the second law. For example, it is impossible to put heat into a system and convert it completely to work, a perfect conversion. The second law forces heat to be rejected (which is wasted), and the heat rejection should be to a different temperature reservoir than the temperature of the reservoir from which the heat is added for the cycle to exist. For refrigerating cycles it also establishes the only possible cycles are the ones in which the COP is less than the ideal COP, which is based only on the two extreme temperatures of the cycle.

    The Carnot cycle can be created by exploiting the constant temperature and constant pressure characteristics of the phase change saturation region of pure substances for constant temperature heat addition and rejection processes. The isentropic processes that link the states between the high and low temperatures can now be processes between high and low pressures within the saturation dome. Movement from a low to high pressure is easy since pumps and compressors are available, and from high to low pressure can be achieved via turbines. This is shown in Fig. 1-4.

    However, in a real device the Carnot cycle shown in Fig. 1-4 is difficult to obtain because of practical considerations. It is difficult to stop process heat addition/removal precisely at state 2. The compression from pressure of 2 to the higher pressure at 3 involves a substance that is mixed liquid and vapor at entry for which reliable compression devices are difficult to make (it is easy to either pump liquid alone or compress vapor). Similar problems occur at the expansion from 4 to 1, perhaps via a turbine, where the inlet is saturated liquid and the outlet is a wet saturated mixture of liquid and vapor. To overcome these problems the standard practical refrigeration cycle is shown in Figs. 1-5a and 1-5b. In this the constant temperature (and constant pressure) heat addition process ends when the entire saturated mixture has turned to saturated vapor at state 2 and a vapor compressor is used from pressure 2 to 3. From 3 to 4 the heat rejection is isobaric, and is only isothermal when in the saturation dome region. The isentropic expansion process from 4 to 1 of the Carnot cycle is replaced by a simple expansion valve that does not preserve entropy but preserves enthalpy. The advantage is that a moving device is avoided and a passive valve is replaced instead.

    Fig 1-5a. Practical refrigeration cycle in (T,s) coordinates

    Fig 1-5b. Practical refrigeration cycle in P-h coordinates

    A pure refrigerant or an azeotropic mixture can be used as the working fluid in the practical refrigerator to maintain constant temperature during phase changes through maintenance at constant pressure. Because of such concerns as high initial cost and increased maintenance requirements, the practical machine has one compressor, and the expander (engine or turbine) is replaced by a simple expansion valve. The valve throttles the refrigerant from high pressure to low pressure. The highest temperature in the cycle is at state 3. Thus, for ideal COP the temperature of 3 has to be used, and not that of state 4. The components of the refrigerator are sketched in Fig. 1-6. Because of the use of a compressor to go from the low temperature state 2 to high temperature state 3, the cycles are called vapor compression cycles.

    Fig 1-6. Components of the refrigerator or air conditioner

    Fundamentals of Fluid Flow

    In addition to density discussed in the previous section, the other property that is of importance in fluid flow is viscosity. The viscosity of the fluid links the shear stress on a fluid layer to the local velocity gradient across the layer. In the terminology of Fig. 1-7, the following relation defines viscosity

    Fig 1-7. Velocity profiles and gradients in viscous flows

    The density of air and water at standard conditions of 68°F and 14.696 psi (sea level atmospheric pressure) are 0.075 lbm/ft³ and 62.3 lbm/ft³, respectively. All fluids are compressible to some degree; fluid density depends on pressure. Steady liquid flow may ordinarily be treated as incompressible, and incompressible flow analysis is satisfactory for gases and vapors at velocities below about 4000 to 8000 fpm, except in very long conduits. The units of viscosity are (force × time)/length². At standard conditions the viscosities of air and water are 1.2×10−5 lbm/ft-s (= 3.7×10−7 lbf-s/ft²) and 6.7×10−4 lbm/ft-s (= 2.1×10−5 lbf-s/ft²), respectively. Kinematic viscosity is viscosity divided by density. The values of kinematic viscosity at standard conditions for air and water are 1.7×10−4 ft²/s and 1.08×10−5 ft²/s, respectively. In general if the Mach number of the fluid is less than 0.30 the fluid can be considered incompressible.

    In this chapter the fluid mechanics of pipe and duct flows are emphasized. This is because in HVAC systems flows in pipes and ducts are dominant methods of fluid transport, both airflow to the conditioned space and liquids that are the working fluids in the systems. External flows, such as flows of large fluid volumes over submerged objects, compressible flows such as in turbines and compressors, and the like are not discussed for the sake of brevity and relevance (or lack thereof).

    Flow in Pipes and Ducts.—The conservation of mass (or continuity equation) follows the same formulation as described before in the thermodynamics section, Equations (8) and (9). For conservation of energy some simplifications and modifications can be made. For fluid mechanics analyses only conservation of mechanical energy is considered. If steady flow is considered along a single path system then the conservation of energy for open systems, Equation (13), can be rewritten as

    Here it is assumed that no work producing shafts or work producing moving boundaries are present and that the conservation of mass then permits the inlet and outlet mass flow rates to be the same. Pressure energy, kinetic energy, and potential energy are the mechanical energies of the flow. Since it was shown in the previous thermodynamics section that for most fluids, especially away from the saturation dome, the internal energy u is primarily a function of temperature, u is not considered a mechanical energy. The conservation of mechanical energy is then stated as

    This is also the Bernoulli’s equation, which is the conservation of mechanical energy in an ideal fluid with no losses, no heat additions/removal, and no work. Bernoulli’s equation shows that the sum of pressure energy, kinetic energy, and potential energy is constant along a streamline of an ideal fluid. Even when temperature variations are present or the internal energies are not constant or heat transfer is present, the above mechanical energy conservation (and modifications presented below in case of losses) can be separately considered before the entire conservation of energy Equation (13) is eventually applied if needed.

    In the case of real fluids, where viscosity introduces drag, loss in mechanical energy occurs. This loss of mechanical energy is usually converted to increase in internal energy or is dissipated as heat, as per general conservation of energy as described in the thermodynamics section. In such a case of losses Equation (26a) of conservation of mechanical energy (presented in the form of energy per unit mass) is rewritten as

    where Eloss = mechanical energy loss per unit mass, or rate of mechanical energy loss per unit mass flow rate.

    If mechanical energy is added to the flow through appropriate forms of work such as shaft work, the above can be modified to include a source term for mechanical energy. This is shown in

    where Emech = mechanical energy added into the pipe or duct flow via a device such as fan, blower, or pump.

    Another consideration in the case of real fluids is the velocity V in Equation (28). The velocity V in the above is the average velocity through the pipe and is obtained from the mass flow rate. Because of the viscosity for real fluids, the velocity varies from zero at the duct or pipe walls to a maximum along the centerline. Since it is not mathematically true that the square of the average velocity is equal to the average of squared velocity unless the velocity profile is flat, this introduces an additional kinetic energy factor α in Equation (27) as follows:

    The kinetic energy factor is the ratio of the true kinetic energy of the flow to the kinetic energy represented by the mean or average velocity. It is shown that the value of this kinetic energy factor is 2.0 for laminar flow in circular pipes and 1.54 for laminar flow in wide rectangular channels. For turbulent flow the value is close to 1.0. Thus, for most HVAC applications, where the flow is turbulent, the original Bernoulli equation Equation (27), with losses is sufficient.

    Reynolds number is an important unitless parameter in fluid mechanics. It is defined for flow in pipes or ducts as

    where D = diameter for circular pipes or a characteristic length for non-circular pipe.

    For other cross-sections it can be the hydraulic diameter Dh which is defined as

    where A = cross-sectional area, and P = perimeter.

    For external flows, characteristic length is usually the length of the object along the direction of the flow, or the value of the length’s local coordinate depending on the application. For pipe and duct flows the flow is laminar if the Reynolds number is less than 2300 and turbulent for greater values. This value of the critical Reynolds number is observed via experiments. For external flows parallel to a surface the flow remains laminar until the local Reynolds number reaches 5 × 10⁵. Thus, the critical Reynolds number for internal flow is 2300 and for the external flow is 5 × 10⁵. If the Reynolds number exceeds the critical value, the flow transitions to turbulence; below the critical value, the flow is laminar.

    Laminar flows are those where viscous effects dominate and the flow is orderly and layered. Fluid particles travel in smooth trajectories, without fluctuations. In steady internal flows, such as in pipes and ducts, the velocity profile is of a form that the forces due to pressure are balanced by viscous shear forces introduced by the presence of a wall (the velocity has to be zero at the wall). This gives rise to parabolic velocity profiles in the pipe or duct, where the velocity vanishes at the walls and is a maximum at the centerline.

    Turbulent flows which typically have higher velocities than laminar flows, involve random perturbations or fluctuations of the velocity and pressure, characterized by an extensive hierarchy of scales or frequencies. Only flows involving random perturbations without order or periodicity are turbulent; the velocity in such a flow varies with time or locale of measurement (Fig. 1-8). Turbulence can be quantified by statistical factors. The velocity most often used in velocity profiles is the temporal average velocity, and the strength of the turbulence is characterized by the root mean square of the instantaneous variation in velocity about the temporal average velocity. The effects of turbulence cause fluid to diffuse momentum, heat, and mass very rapidly across the flow. Because of the rapid fluctuations, fluid particles do not travel in smooth trajectories. The fluctuation allows particles to cross layers and thus cause a greater uniformity of flow properties and characteristics than in the laminar case. Because of this the turbulent velocity profiles in pipes or ducts are flatter throughout the core of the flow around the centerline and only fall off to zero velocity at the walls in a small region near the walls. It is because of the flatter velocity profile that the kinetic energy factor in Equation (29) is near unity, usually ranging from 1.01 to 1.10.

    Fig 1-8. Velocity fluctuations with respect to time in turbulent flows

    Time average velocity is defined by the Equation (32)

    The laminar and turbulent velocity profiles in a pipe flow are schematically compared in Fig. 1-9. Turbulent flow profiles are flat compared to the more pointed profiles of laminar flow. Near the wall, velocities of the turbulent profile must drop to zero more rapidly than those of the laminar profile, so the shear stress and friction are much greater in the turbulent flow case.

    The velocity profiles shown in Fig. 1-9 correspond to fully developed flows. Fully developed flow regions are far away from inlets, from sections of sudden changes in cross-section, and from sources of mechanical energy input. The entrance flow region is the region from an inlet to the location where the fully developed region starts. It corresponds to the length Le of the pipe or duct needed for the flow to gradually change from its inlet profile to the new conditions.

    Fig 1-9. Laminar and turbulent velocity profiles in a circular duct

    Note that if the cross-sectional area remains constant then by the conservation of mass the average velocity throughout the pipe, including the inlet and fully developed regions, remains the same although the shape of the profile changes.

    With laminar flow following a rounded entrance, the entrance length Le depends on the Reynolds number:

    At Re = 2000, a length of 120 diameters is needed to establish the fully developed parabolic velocity profile. However, the pressure gradient reaches the developed value of much sooner. With turbulent flow, a length of 80 to 100 diameters following the rounded entrance are needed for the velocity profile to become fully developed, but the friction loss per unit length reaches a value close to that of the fully developed flow value more quickly. After six diameters, the loss rate at a Reynolds number of 10⁵ is only 14% above that of fully developed flow in the same length, while at 10⁷ it is only 10% higher. For a sharp entrance, the flow separation causes a greater disturbance, but fully developed flow is achieved in about half the length required for a rounded entrance. With sudden expansion, the pressure change settles out in about eight times the diameter change (D2−D1), while the velocity profile takes at least a 50% greater distance to return to fully developed pipe flow.

    The mechanical energy loss in a duct or pipe of constant cross-section due to friction is given by the following:

    where f = friction factor.

    Fig 1-10. Friction factor (Moody’s chart)

    Losses due to friction are traditionally termed major losses in a pipe or duct system because for long pipes this loss dominates. However, for HVAC applications this may not be the case, and so the terminology of major losses may not be as appropriate as simple friction loss. Here L is the length of the pipe and D is the diameter. For noncircular ducts, the diameter may be replaced by the hydraulic diameter Dh. Friction factor is a function of the Reynolds number and the relative roughness of the pipe or duct walls. For large Reynolds numbers its value is fairly constant and is only a function of the roughness ε/D, where ε is the average height of roughness. This region is called the fully rough region. The value of friction factor is obtained via experiments and the values are given in Fig. 1-10 and in Equations (35) to (38).

    In the laminar region, where the Reynolds number is less than 2300, the friction factor is independent of the roughness. This is because the dominant viscous effects suppress any fluctuations introduced by roughness. The value of f for this region is given as

    for circular cross-section pipes. In non circular pipes exact values for f can be derived for laminar flow, and the use of hydraulic diameter in Equation (35) may lead to significant errors.

    In turbulent flows for smooth pipes the friction factor is empirically correlated as

    Another correlation for smooth pipes is

    For pipes and ducts that are not smooth, where a wall roughness ε is defined (in the units of length), the relative roughness ε/D plays an important role. In the fully rough region where the influence of Reynolds number is very small, the turbulent friction factor has been empirically correlated by

    In the other turbulent regime between the smooth walled pipes and the fully rough region several empirical correlations exist. A commonly used correlation is the Colebrook function given by

    The use of hydraulic diameter Dh in the above correlations for turbulent flows is quite acceptable since errors up to 5% may be introduced as observed experimentally.

    Mechanical energy losses other than friction loss described above in a pipe or duct are usually correlated by using a loss coefficient K:

    For inlets, the V in Equation (41) is velocity in the pipe after the entrance, and for outlets it is the velocity in the pipe before the exit. Thus, which velocity to use is a function of the geometry or of the type of the fitting or fixture. For a sudden expansion it is velocity before expansion and for a sudden contraction of cross sectional area it is velocity after contraction. The values of K are compiled via experiments and will be discussed in later chapters. It may be noted that although the true loss of mechanical energy occurs over a distance the formula lumps its effects at one location. For example, the inlet loss occurs over the entrance region but is lumped at the entrance by this formula. Note that these effects do not include the friction loss over the corresponding length, but only the additional losses associated with disturbances of flow and transitions. For systems with long pipe lengths these losses are a small portion of the total losses where the friction losses dominate and are traditionally called minor losses. However, for HVAC systems this is not the case and the multiple bends and junctions, size changes, and the presence of various fixtures and fittings ensure that these effects are a very significant part of the total loss.

    Some Details of Relevant Flow Fields.—The presence of walls in fluid flows introduces velocity gradients in the flow field since the fluid necessarily has to be at zero velocity at the wall. This effect, termed friction, usually is present in the form of a boundary layer in fast moving flows. A boundary layer is the slender region near the wall where the velocity goes from the zero wall velocity to the free stream velocity. All the viscous effects are concentrated in the boundary layer. For flow around bodies, this layer (which is quite thin relative to distances in the flow direction) encompasses all viscous or turbulent actions, causing the velocity in it to vary rapidly from zero at the wall to that of the outer flow at its edge. Boundary layers are generally laminar near the start of their formation but may become turbulent downstream of the transition point.

    For conduit or pipe flows, spacing between adjacent walls or within the pipe diameter is generally small compared with distances in the flow direction. As a result, layers from the walls meet at the centerline to fill the conduit. The region after the layers meet is the fully developed flow region that was discussed earlier. The length of the pipe or conduit where the boundary layers are still growing and have not yet met at the centerline is called the entrance region. Near the start of the straight conduit or pipe, the layer is very thin (and laminar in all probability), so the uniform velocity core outside has a velocity only slightly greater than the average velocity. As the layer grows in thickness, the slower velocity near the wall needs to increase in the uniform core to satisfy continuity. As the flow proceeds, the wall layers grow (and the centerline velocity increases) until they join, after an entrance length Le. Application of the Bernoulli relation to the core flow indicates a decrease in pressure along the layer.

    However, if the cross-sectional area is increasing along the flow so that average velocity decreases along the flow, the adverse pressure gradient can lead to flow separation. The development of the boundary layer in an adverse-pressure gradient situation (velocity at edge of layer decreasing in flow direction) causes the separation of the boundary layer where downstream from the separation point the fluid backflows near the wall. Separation is due to flow near the wall no longer having energy to move into the higher pressure imposed by the decrease in velocity at the edge of the layer. The locale of this separation is difficult to predict, especially for the turbulent boundary layer. Analyses verify the experimental observation that a turbulent boundary layer is less subject to separation than a laminar one because the flow in the turbulent layer has greater kinetic energy.

    Fig 1-11a. Flow through orifice

    Fig 1-11b. Flow through sudden contraction or expansion

    Flow separation is observed in several situations in HVAC equipment. Examples of flow through an orifice and through contractions and expansions are shown in Figs. 1-11a and 1-11b. The manifestation of separation is sudden pressure drop that is not recovered in the flow.

    In liquid flow, gas- or vapor-filled pockets can occur if the absolute pressure is reduced to vapor pressure or less. In this case, a cavity or series of cavities form. This lowering of pressure may be due to flow separation or poor selection of operating parameters or equipment design. This is called cavitation and initial evidence of cavitation is the collapse noise of many small bubbles as they are carried by the flow into regions of higher pressure. The severity of cavitation increases as velocity increases or pressure decreases. Collapse of the cavities on or near solid boundaries becomes so frequent that the cumulative impact in time results in damage in the form of cavitations erosion of the surface or excessive vibration. As a result, pumps can lose efficiency or their parts may erode locally.

    Noise from Fluid Flow.—Noise from fluid flow is especially important in HVAC systems where, for example, noisy ducts inside buildings can significantly cause complaints. Noise in flowing fluids results from unsteady flow fields and can be at discrete frequencies or broadly distributed over the audible range. With liquid flow, cavitation results in noise through the collapse of vapor bubbles. Noise in pumps or fittings (such as valves) can be easily eliminated by raising the system pressure. With severe cavitation, the resulting unsteady flow can produce by-product noise from induced vibration of adjacent parts.

    Noise produced in pipes and ducts is especially associated with the loss through the valves and fittings. The sound pressure of noise in water pipe flow increases linearly with the head loss; the broadband noise increases, but only in the lower frequency range. Fitting-produced noise levels also increase with fitting loss (even without cavitation) and significantly exceed noise levels pipe flow. The relation between noise and loss associated with is because both involve excessive flow perturbations. A valve’s pressure flow characteristics and structural elasticity may be such that for some operating point it oscillates, perhaps in resonance with part of the piping system, to produce excessive noise. A change in operating point conditions or details of valve geometry can result in significant noise reduction.

    Pumps and blowers are strong potential noise sources. Turbo machinery noise is associated with blade-flow occurrences. Broadband noise results from vortex and turbulence interaction with walls and is primarily a function of the operating point of the machine. For blowers, noise is a minimum at the peak efficiency point. Narrowband noise also occurs at the blade, crossing frequency and its harmonics. To reduce this noise, a designer increases the clearances between impeller and housing, and spaces impeller blades unevenly around the circumference. However, this may be impractical because they lead to other problems.

    Fundamentals of Heat Transfer

    Heat is transfer of energy due to temperature difference. Solution of conservation of energy Equations (11) to (13) require the accurate assessment of heat transported into a closed or open system. Heat transfer theory is needed to quantify the rate at which heat will transfer for a given set of conditions. Although application of thermodynamic law requires the value of heat transfer be known, it does not provide any method to ascertain whether the required or prescribed heat transfer rate is feasible for the given system and available temperature differences. Thermal energy transfer always occurs in the direction of decreasing temperature.

    Thermal conduction is the mechanism of heat transfer whereby energy is transported between parts of a continuum by the transfer of kinetic energy between particles or groups of particles at the atomic and subatomic levels. In liquids and gases, conduction is caused by elastic collision of molecules; in liquids and electrically nonconducting solids, it is caused by oscillations of the lattice structure (or phonons); in metals it occurs due to the motion of free electrons. In the study of engineering heat transfer, in contrast with physics, the underlying microscopic mechanism is not of interest and the entire conduction effect is lumped into understanding thermophysical properties.

    Thermal conductivity k is defined as the ratio of heat flow rate per unit area to the local temperature gradient. This is given by Fourier’s law,

    In Equation (42) steady heat flow along the x direction is considered. The units for k are Btu-ft/hr-ft²-°F and units of temperature T and distance x are °F and ft, respectively. The minus sign indicates that heat flow is positive in the direction of decreasing temperature (i.e., negative gradient). Thermal conductivity is an intrinsic property of a material and is tabulated in handbooks and property tables.

    In solid opaque bodies, thermal conduction is the most significant heat transfer mechanism because there is no flow of material, and transport by other heat transfer mechanisms such as convection and radiation are not present. In fluids, thermal conduction dominates in the region very close to the flows solid boundary. However, in fluids the effects of conduction near the boundary are lumped into heat convection coefficients.

    The distinction made between thermal convection and thermal conduction is this: Heat convection involves energy transfer or exchange at an interface, notably, the solid-fluid interface— the walls. Heat conduction occurs throughout the entire body of the transport mechanism. The heat transfer coefficient h is defined at the interface in the fluid region so that the rate of heat flow is related to the temperature difference by

    Equation (43) relates the heat flow rate from the surface to the fluid. The units for h are Btu/hr-ft²-°F. For pipe flows the fluid temperature in Equation (43) is the bulk mean temperature at the cross-section where the convective heat transport is being considered.

    The heat transfer coefficient is an engineering concept that combines all microscopic energy transport processes near a solid-fluid interface into a single coefficient. The heat transfer coefficient is strongly a function of flow conditions and increases as flow velocities increase. It is not an intrinsic property of any material and it must computed or measured at the required flow configurations and geometries.

    Thermal convection can be classified into two types: forced and natural. Forced convection occurs when fluid is flowing due to external mechanisms such as the case of it being driven by fans or pumps. Natural convection occurs in fluid that is normally quiescent, but is forced into motion by thermal buoyancy effects. For example, hot fluid near a hot wall rises because its density is lowered and colder fluid moves in to replace it, thereby setting up a loop that removes heat from the hot wall. In this case the value of h is strongly a function of temperature, the variations of the density with temperature, and the geometry of the system.

    For internal flows through pipes and ducts the convective heat transfer coefficient stays constant along the direction of the flow in the fully developed region, i.e., after the entrance length region. For external flows the local convective heat transfer coefficient changes along the direction of the flow because the boundary layer keeps on growing, as opposed to internal flow where the boundary layers meet at the centerline at the end of the entrance region and no further changes in flow occur. The forced flow heat transfer correlations can be written generally as

    Nusselt number is given by

    The length in the Reynolds number is the same characteristic length as in the Nusselt number. For pipes the characteristic length is the diameter D or the hydraulic diameter Dh for non-circular pipes. For flat surfaces the length to be used is the entire length along the direction of flow for computing average heat transfer coefficient, or the local coordinate along the direction of the flow to compute the local heat transfer coefficient. The Prandtl number is a ratio of viscous and thermal diffusivities of the fluid and given by

    where α = thermal diffusivity.

    The thermal diffusivity is given by

    The general forced convection Equation (44) remains valid for flows inside pipe and ducts, as well as over flat surfaces and other objects. The constants C, n, and m for the above forced convection correlation are given in Table 1-1.

    Table 1-1. Forced Convection Factors

    a Note: Reynolds number in this case uses the maximum velocity in the space between tubes.

    In the case of entry length flow in circular pipes the laminar correlation is modified as follows for the constant surface temperature case:

    The Reynolds and Nusselt number are based on the diameter. The effects of thermal entry length on turbulent flows are not significant and the fully developed case can be used. Similarly, the effects of surface roughness are not noticeable in the heat transfer coefficient for turbulent flows (in contrast to the fluid mechanics friction factor), and thus need to be considered.

    For flows parallel to flat surfaces the initial boundary layer is always laminar and makes the transition to the turbulent layer after the local Reynolds number becomes greater than critical. Thus, for flows that do become turbulent, the average heat transfer coefficient involves the initial laminar correlations, and then the turbulent. The average heat transfer coefficient for such cases is

    Here the characteristic length is the length of the surface L in the direction of the flow, provided that Re based on this length is greater than the critical value 5 × 10⁵.

    A general formula for the average heat transfer coefficient for the external cross-flow over a cylinder, over the entire range of Reynolds numbers, is given by

    The general correlation for natural convection is given by

    where C, n = are constants whose values depend on the operating range and geometry.

    The Grashof number Gr is given by

    where ΔT = magnitude of the difference between the surface temperature and the free-stream or bulk fluid temperature.

    The characteristic length L in the above natural convection correlations represents diameter for horizontal circular pipes, height for vertical plates or pipes, and radius for spheres. For vertical flat surfaces, the natural convection flows remain laminar until the value of (GrPr) reaches 10⁹. The Rayleigh number Ra is product of the Grashof and Prandtl numbers, and thus the critical Rayleigh number for natural convection over vertical flat plates is 10⁹.

    The values of the constants in Equation (51) the correlation for natural convection are given in the following table:

    Table 1-2. Natural Convection Factors

    A correlation that covers the entire range of Ra for a vertical plate has been developed as

    Similarly a general correlation for a horizontal cylinder or pipe over the entire operating range for Ra < 10¹² has been formulated as

    Thermal radiation is the transport of heat via electromagnetic waves or photons. Whereas conduction and convection require a medium, radiation transport can occur in a vacuum. The rate of thermal radiant energy emitted by a surface depends on its absolute temperature. A surface is called black if it can absorb all incident radiation. The total energy Wb emitted per unit time per unit area of black surface to the hemispherical region above it is given by the Stefan-Boltzmann law:

    Substances and surfaces diverge in various ways from the Stefan-Boltzmann laws. Wb is the maximum emissive power at a surface temperature. Actual surfaces emit and absorb less than these maximums and are called nonblack. The emissive power of a nonblack surface at temperature T radiating to the hemispherical region above it is written as

    where  ε = hemispherical emittance.

    Emittance is a function of the material, the condition of its surface, and the temperature of the surface.

    When radiation energy is incident on a surface, such as solar flux incident on a wall, the fraction of energy absorbed by the surface is the radiation absorptance. For engineering applications absorptance is taken to be equal to emittance. The rest of the incident radiation is reflected back and the reflectance is equal to 1 – absorptance, which is equal to 1 – emittance.

    If the surface is not completely opaque then some of energy is transmitted. An example is a pane of glass in a window. Incident radiation energy is either absorbed, reflected or transmitted. The relationship between these is

    Overall Heat Transfer.—Many applications in HVAC involve different modes of heat transfer occurring in series or in parallel. For example, the loss of heat from a room to the cold outside ambient space through the walls involves three heat transfer steps: the transfer of heat from the indoor air to the inner surface of the wall via convection, the transfer of heat from the inner surface of the wall to the outer via conduction through the wall; and the transfer of heat from the outer surface to the cold ambient outdoors. If the temperatures of the indoor air and outdoor ambient spaces are T1 and T2 respectively, the rate of heat transport is given by

    where U = overall heat transfer coefficient.

    If the wall temperatures are Tw1 and Tw2, respectively at the inner and outer surfaces, and the corresponding heat transfer coefficients are h1 and h2, then the same heat transfer rate evaluated through the three individual steps is given by

    Combining Equation (57) and Equation (58) gives the expression for the overall heat transfer coefficient for this case as

    The overall heat transfer coefficient can also be evaluated by considering thermal resistances. Through an analogy of flow of heat to flow of electricity through conductors where the driving potential is temperature difference, the thermal resistance can be defined as

    Sometimes Rth is written without subscript as R. When many thermal steps are in series the thermal resistances add in series and the total resistance is related to the overall heat transfer coefficient. This is elaborated in Table 1-3.

    Thus, the following can be used to compute the overall heat transfer coefficient when the resistances are in series:

    If the resistances are not completely in series, then Equation (61) can be modified appropriately to account for the correct series/parallel addition of the thermal resistances.

    Fins and Extended Surfaces.—Heat transfer from a surface, such as indoor radiator surfaces in heating and cooling systems, can be increased by attaching fins or extended surfaces that increase the area available for convective heat transfer. Fins provide a more compact heat transfer device with lower material costs for a given performance. Since heat flow rate is directly proportional to heat transfer coefficient, surface area, and temperature difference, increasing any of these three parameters will increase heat flow rate. However, increasing the convective heat transfer requires active processes such as introducing fans. Increasing surface area via extended surfaces is a passive, inexpensive method of increasing heat flow.

    Table 1-3. Solutions for Some Steady State Thermal Conduction Problems

    Fin efficiency ϕ is defined as the ratio of the actual heat transferred from the fin to the heat that would be transferred if the entire fin were at its root or base temperature. This measure is needed because the temperature of the fin varies from its base to the tip as heat flows from the base to the tip and heat is lost from the lateral surfaces. Fin efficiency is low for long fins, thin fins, or fins made of low thermal conductivity material. For an infinitely long fin the fin efficiency asymptotically reaches zero. Fin efficiency decreases as the heat transfer coefficient increases because of increased heat flow. For natural convection in air cooled condensers and evaporators, where h for the air side is low, fins can be fairly large and fabricated from low conductivity materials such as steel, instead of copper or aluminum. For condensing and boiling, where large heat transfer coefficients are relevant, fins must be very short for optimal use of material. The value of fin efficiency is a function of fin geometry, and some shapes are shown in Fig. 1-12.

    Fin effectiveness ε is a more important concept than ϕ, since it relates the heat flow from a surface Ab if it is covered by a fin to heat flow from a surface and not covered by a fin. It is defined in relation to fin efficiency as

    Fig 1-12. Efficiency of straight fins

    As the physical definition indicates effectiveness is greater than unity. It should be significantly greater than one to justify the additional expense and space needed for the fins. As fin length increases effectiveness will increase asymptotically to a maximum. Here Tb is the surface temperature of the unfinned surface, which is also equal to the temperature of the base of the fin. The maximum value of fin effectiveness for an infinitely long straight fin is given by

    where yb = half the thickness of the fin.

    If a surface A is covered partially by a fin (or fins) so that the area covered by fin(s) is Ab and that not covered is (A – Ab), then the heat flow from the entire area is given by

    The surface effectiveness εs of fins for the entire area A that is partially covered by the fins, as above, is given by

    Thus, the surface effectiveness with fins is always greater than one.

    Fig 1-13. Fin efficiency for annular fins of constant thickness

    Some Details of Heat Exchange.—In this section some mathematical developments of heat transfer in flowing streams are

    Enjoying the preview?
    Page 1 of 1