Sei sulla pagina 1di 36

11

Thermodynamic Modeling and Analysis

The main objectives of this chapter are twofold: (1) to present and discuss important factors that affect heat exchanger performance, and (2) to introduce a basic analysis for the thermodynamic design and optimization of heat exchangers. A quest for answers regarding the first objective will help us to identify the important factors that affect heat exchanger effectiveness, to quantify the effects of these factors, and to provide guidelines for a qualitative assessment of the effectivenesses of the exchangers with different flow arrangements but with a given, identical design task. The second objective is to define a figure of merit for assessing the thermodynamic efficiency of a heat exchanger and to present an approach to thermoeconomic considerations. In Section 11.1, the differences between a heat exchanger as a component and as part of a system are identified. In Section 11.2, a detailed modeling of a heat exchanger using energy balances only (i.e., the first law of thermodynamics) is provided for the determination of heat exchanger effectiveness and temperature distributions. In Section 11.3, a combined approach based on both the first and second laws of thermodynamics is introduced to quantify inherent irreversibilities in a heat exchanger. The most important source of irreversibility is heat transfer across the finite temperature differences, which is discussed first. Fluid mixing and fluid friction, as additional sources of irreversibility, are studied next. A temperature cross phenomenon is then discussed in detail in Section 11.4 by evaluating entropy generation in a 1-2 TEMA J shell-and-tube heat exchanger. Using all the analysis tools presented in the first four sections, a heuristic approach to an assessment of heat exchanger effectiveness is developed in Section 1 1.5. In Section 1 I .6, energy, exergy, and cost balances important for analysis and optimization of heat exchangers are presented. Finally, a thermodynamic criterion for evaluation/selection of heat transfer surfaces is summarized in Section 11.7.

11.1 INTRODUCTION

Traditionally, modeling of a heat exchanger is based on energy balances (i.e., on the consequences of both the first law of thermodynamics and the mass conservation principle), so only the concepts of heat transfer rate and enthalpy rate change would
135

736

THERMODYNAMIC MODELING AND ANALYSIS

suffice for such an analysis.+ For an adiabatic heat exchanger (see the assumptions in Section 3.2. l), the enthalpy rate change of one fluid stream must be equal to the enthalpy rate change of the other, being at the same time equal to the exchanger heat transfer rate. This simple energy balance statement will be used in subsequent sections in the differential form to model spatial distributions of temperatures of both fluid streams. In Chapter 3, we also used the energy balances (both differential and overall), but only to determine the heat exchanger effectiveness without determining the temperature distributions. Let us start with a review of the analysis of heat exchanger design methods discussed in Chapter 3 that relies on a relationship that can be presented in generalized form as follows: effectiveness/) rate q factor

heat capacity rate or thermal conductance

(temperature)
difference

&Cmin AT,,,

in E-NTU method in P-NTU method in LMTD method in @-NTU method (11.1)

P,CI AT,,,
FUA AT,, @UA AT,,,

Equation (1 1.1) is based on energy balances formulated as consequences of the first law of thermodynamics. It is important to note that each of the methods implied by Eq. (1 1.1) assumes the determination of either an effectiveness factor (heat exchanger effectiveness E or temperature effectiveness PI)or a correction factor ( F o r @)as a function of design parameters, e.g., N T U and C*. We use the expressionfactor as a generic term to indicate a common first law of thermodynamics origin for both the effectiveness and correction factors. Of course, the correction factors do not have the same physical meaning as the effectiveness factors. Each relationship in Eq. (1 1.1) involves a temperature difference, either the maximum imposed temperature difference AT,,, or the logarithmic-mean temperature difference ATl,,,. It has been demonstrated in Sections 3.3 and 3.5 that relationships between effectiveness factors and the pertinent design parameters can be devised from the heat transfer model of a heat exchanger. In some cases, these relationships can even be obtained without a detailed study of internal heat transfer interactions. Moreover, a designer who already has these relationships would be able to calculate the effectiveness for the given set of parameters and execute a design method procedure without a need to study temperature distributions of a selected flow arrangement as outlined in Sections 3.9 and 9.2 through 9.5. So one would treat the heat exchanger as a black box for determination of the overall heat transfer surface area or heat transfer performance of an exchanger as a component. In such a case, there is no need to know temperature distributions. For example, for any exchanger (for which the
In our analysis, the enthalpy rate change is the rate change of enthalpy of a fluid stream caused by heat transfer interaction between the two fluid streams. From the thermodynamic point of view, note that the heat transfer rate is an energy interaction (not a change in the fluid property), while the enthalpy rate change represents a change of fluid property caused by existing heat interaction. This distinction is important for understanding of the interpretation of heat exchanger performance and will be emphasized in more detail later through the introduction of several advanced concepts of thermodynamics.

INTRODUCTION

737

effectiveness relationships are already known or can be determined using the matrix formalism mentioned in Section 3.1 1.4), an engineer needs only a relationship between the effectiveness/correction factor and design parameters, without detailed insight into local temperature distributions. Thus, a misleading conclusion may be reached: that the only information a designer should possess concerning a flow arrangement is the relationship between an effectiveness factor and design parameters (e.g., PI -NTU1,E-NTU, or F-P relationship). Analysis presented so far does suffice for a design procedure for a heat exchanger with an already defined effectiveness relationship. However, a very important and still unanswered question should be addressed as well. Why does an effectiveness factor (say, heat exchanger effectiveness) have a high (or low) value for a given flow arrangement (especially for a complex one) compared to the corresponding value for another flow arrangement (for the same set of design parameters)? For example, we do know that a crossflow heat exchanger has less exchanger effectiveness than for a counterflow exchanger (for the same set of design parameters NTU and C*). The only rational explanation that we can offer at this point (in addition to the intuitive ones) is that the effectiveness relationship for a crossflow exchanger simply provides a smaller numerical value for E or P for the given heat capacity rate ratio and NTU than does E or P for a counterflow exchanger. In addition, As/ANTU for a fixed heat capacity rate ratio C* is different for counterflow than for crossflow, for NTU,i, < NTU < co.For NTU 2 4, this gradient is almost identical. For NTUmi, 5 0.4, all flow arrangements provide almost identical effectiveness values for a given set of design parameters [see Eq. (3.89)]. Why is that so? The reasoning will become clear when we present an astonishingly simple heuristic approach based on the second law analysis of exchanger flow arrangements. Also, we present a thermodynamic performance figure of merit, the efficiency of a heat exchanger from a system viewpoint. Consequently, these analysis tools will help us in assessing relative magnitudes of exchanger effectiveness for complex flow arrangements for the selection of an appropriate flow arrangement for a specified task. This understanding will also become valuable in finding an optimum heat exchanger design from a system viewpoint.

11.1.1 Heat Exchanger as Part of a System

Heat exchangers in numerous engineering applications are only one of many components of a system. Thus, the design of a heat exchanger is inevitably influenced by system requirements and should be based on system optimization rather than component optimization. An objective function for such system-based optimization is influenced by the main features of heat exchanger operation. For a given set of input data (e.g., flow rates and inlet temperatures), exchanger geometry, and other pertinent information, the output data (e.g., the outlet temperatures) will depend on heat transfer and fluid flow phenomena that take place within the boundaries of the heat exchanger. So even though one seeks a system optimum, in the process of determining that optimum, one must fully understand the features of the exchanger as a component. Since heat exchangers are used in many systems, we do not attempt any specific system analysis or process integration. We discuss only the basic thermodynamic aspects. Despite exact mathematical/numerical results obtained through system-based optimization, the designer should know that the heat exchanger design (sizing) problem studied is a complex problem that has no single exact solution at all. In all but trivial cases, a

738

THERMODYNAMIC MODELING AND ANALYSIS

designer must deal with uncertainty margins of the input data, in addition to numerous assumptions. Usually, a range of data (say, for a cost analysis in an optimization routine), and not a single set of parameters, must be considered. As shown in Fig. 2.1, for every case considered in the top left box of the problem specification, one arrives at an optimum solution at the end of the process of Fig. 2.1. Hence, there can be many (and not only one) optimum solution for a given exchanger sizing problem as provided by different heat exchanger manufacturers. With that in mind, the reader should understand the limitations of results obtained by modern computer software for design, optimization, and system integration.
11.1.2 Heat Exchanger as a Component

Before a system-based optimization can be carried out, a good understanding of the exchanger as a component must be gained. In addition to rating or sizing, this may include information about temperature distributions, local temperature differences, hot and cold spots, pressure drops, and sources of local irreversibilities-all as functions of possible changes of design and/or process variables and/or parameters. The outlet state variables of the fluids depend on the efficiency of the heat transfer process influenced by fluid flow and heat transfer phenomena within the exchanger. A measure of this efficiency is not defined exclusively by heat exchanger or temperature effectivenesses because it gives relevant but limited information about heat exchanger performance since the influence of irreversibility, as discussed in Section 11.3 is not included. Thus, the key questions to be answered involve how to define exchanger efficiency, and how the heat transfer and fluid flow processes (manifested within the heat exchanger boundaries) affect the exchanger effectiveness and thermodynamic efficiency (see Section 11.6.5). To answer these questions, we first identify the important heat transfer/fluid flow phenomena in the operation of a heat exchanger having an arbitrary flow arrangement in Section 11.3. Design of a heat exchanger as a component is to a large extent an engineering art. So, despite high sophistication in heat exchanger thermal modeling, some of the final decisions (in particular those related to optimization) are based on qualitative judgments due to nonquantifiable variables associated with exchanger manufacturing and other evaluation criteria. Still, analytical modeling-a very valuable tool-is crucial to understanding the relevant thermal-hydraulic phenomena and design options and various venues for design improvements. In structuring this chapter, special attention is devoted to a balanced use of both rigorous mathematical modeling (Sections 11.2 through 11.4) and qualitative analysis and heuristic judgments (Section 11.5). The results based on mathematical modeling, although elegant and transparent concerning the influences involved, always carry within them all consequences of numerous assumptions and often simplifications. So the primary purpose of our study in this chapter is to gain a good understanding of the factors that affect exchanger performance, not necessarily to provide new tools for design and system-based optimization of a heat exchanger.

11.2 MODELING A HEAT EXCHANGER BASED ON THE FIRST LAW OF THERMODYNAMICS

An important objective of the material presented in this section is to learn how to model a heat exchanger to determine temperature distributions. In Chapter 3, we focused on the

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

739

determination of the exchanger efficiency factor E , P, and F [see Eq. (1 1. l)], through various heat exchanger basic design methods; we did not pay any particular attention to fluid temperature distributions and their relationship to exchanger performance. Let us now consider the distribution of local temperatures and temperature differences in a heat exchanger having simple counterflow and parallelflow arrangements. Both these arrangements correspond to two limiting cases of the same geometrical situation: The two fluid streams are flowing in geometrically parallel orientation but in opposite or same directions to each other, thus providing the largest and smallest heat exchanger effectiveness values (see Figs. 3.1 and 3.8 and Table 3.3). It is assumed, by definition, that both fluids change their respective temperatures only in flow directions (i.e., the local temperature distribution is uniform for a fluid across a flow cross section). Subsequently, we consider a more complex situation with a cross flow of working fluids and the possibility of local mixing along the flow direction.

11.2.1 Temperature Distributions in Counterflow and Parallelflow Exchangers


In Fig. 11.1 a schematic of a counterflow heat exchanger is presented. The assumptions formulated in Section 3.2.1 are invoked here. In general, flow directions may be either in the positive (in Fig. 1 1.1, from left to right) or negative direction in relation to the axial coordinate (i.e., counterflow/parallelflow). The energy balances for the respective control volumes can be written as follows using the first law of thermodynamics and following rigorously the standard sign convention for heat/enthalpy rate flows across the control volume boundary (positive if entering into the system and negative if leaving the system). For fluid 1 only (the elementary control volume in Fig. 1 1 . 1 ~ ) :

(4

(b)

(d

FIGURE 11.1 Energy balance control volumes for a counterflow arrangement. The control volumes are represented by rectangular areas. (a) Control volume for fluid 1 differential energy balance; (b)control volume for fluid 2 differential energy balance; (c) control volume for both fluids 1 and 2 differential energy balance. Enthalpy rates H, = (ricp),q,j = 1 or 2 .

740

THERMODYNAMIC MODELING AND ANALYSIS

where i l = $1 or -1, for the same or opposite (positive or negative) direction of fluid 1 with respect to the positive direction of the x axis, respectively. For fluid 2 only (the elementary control volume in Fig. 1l.lb):

where i2 = +1 or - 1, for the same or opposite (positive or negative) direction of fluid 2 with respect to the positive direction of the x axis, respectively. For both fluids 1 and 2 (the elementary control volume in Fig. 11.1~):

Note that an assumption of uniform distribution of the total heat transfer surface area A along the flow length L means that dA/dx = AIL; that is, U dA = ( U A / L )dx. Let us, without restricting the model generality, fix the direction of fluid 1 to be in a positive axial direction (il = + I ) while i2 = i l [i.e., i2 = - 1 for counterflow (see Fig. 1 l.l), or i2 = +1 for parallelflow]). Rearranging Eqs. (1 1.2)-( 1 1.4), we obtain
dT1 UA 2 - TI) L dx dT2 UA i2(hcp)2 -= - ( T I - T2) dx L
(rizc) - = - ( T

(11.5) (11.6) (11.7)

Note that only two of the three balance equations are sufficient to define the two temperature distributions. For example, either Eqs. (1 1.5) and (1 1.6) or Eqs. (1 1.5) and (1 1.7) can be utilized. Subsequently, distribution of the temperature difference along the heat exchanger can be determined. To close the problem formulation, a set of boundary conditions is required at the heat exchanger terminal points. For a parallelflow exchanger, the inlet temperatures for both fluids are known at x = 0. For the counterflow exchanger, the known inlet temperatures are on the opposite sides of the exchanger, at x = 0 and x = L, respectively. Explicitly, these conditions are
x = 0 for parallelflow x = L for counterflow

TI =

at x = 0

T2 = T2,; at

(11.8)

Equations (1 l S F ( 11.8) are made dimensionless with the following variables: (11.9)

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

741

and design parameters NTUl and R I , as defined by Eqs. (3.101) and (3.105). Hence, (11.10)

(1 1.1 1) ( 1 1.12)
The boundary conditions are as follows:
01=0
Oz= 1

atE=O

(11.13) (11.14)

at

E = 0 for parallelflow E = 1 for counterflow

The set of relationships given by Eqs. (1 1.5)-( 11.8) or (1 l . l O t ( 1 1.14) define a mathematical model of the heat transfer process under consideration in terms of temperature distributions for both fluids. For example, one can solve Eqs. (1 1S), (1 1.7), and (1 1.8) to obtain temperature distributions (as presented in Figs. 1.50 and 1.52). This can be done for virtually any combination of design parameters (NTUI and R 1 )for both parallelflow and counterflow arrangements without a need for separate mathematical models. Some mathematical aspects of the solution procedure and thermodynamic interpretation of the results will be addressed in the examples that follow. A rigorous and unified solution of the parallelflow heat exchanger problem defined above is provided in Example 11.1 (SekuliC, 2000). The relation between the heat exchanger and/or temperature effectiveness as a dimensionless outlet temperature of one of the fluids and a thermodynamic interpretation of these figures of merit is emphasized in Example 1 1.2. An approach to modeling more complex situations, such as a 1-2 TEMA J shell-and-tube heat exchanger or various crossflow arrangements, is left for an individual exercise (see Problems 1 1.1 through 1 1.10 at the end of the chapter).
Example 22.1 Determine temperature distributions of two parallel fluid streams in thermal contact. The fluid streams have constant mass flow rates and constant but different inlet temperatures. Show that a uniJied solution procedure can be formulated for both parallelflow and counterflow arrangements.

SOLUTION
Problem Data and Schematic: The two fluid streams flow in a parallel geometric orientation as presented in Fig. El l. 1A. Both parallelflow and counterflow arrangements are considered (i.e., fluid 2 can be in either one of two opposite directions). Inlet temperatures, mass flow rates, and fluid properties are known. All the geometric characteristics of the flow passages are defined as well. Determine: The local temperatures of both fluids as functions of the axial distance along the fluid flow direction.

742

THERMODYNAMIC MODELING AND ANALYSIS

Fluid 1

Fluid 2 (parallelflow)

x = L (counterflow)

5 =xlL

5-1

ip=-1

FIGURE El 1.1A Fluid flow orientations in counterflow and parallelflow heat exchangers.

Assumptions: It is assumed that thermal interaction between the fluids takes place under the assumptions described in Section 3.2.1 Analysis: Any two of three differential balances given by Eqs. (1 1. lo)-( 1 1.12) together with the boundary conditions given by Eqs. (1 1.13) and (1 I . 14) describe the theoretical model for analysis. Let us define the model of this heat transfer process using Eqs. (1 1.10) and (1 1.12)-( 1 1.14). A general solution will be obtained utilizing the Laplace transforms method (Sekulib, 2000), although several other methods can be used as well (see Section 3.1 1). The rationale for using this particular method is that it can be applied efficiently to a number of more complex situations, such as for a crossflow arrangement with both fluids unmixed (see Problem 11.2). Applying Laplace transforms to Eq. (1 1.10) yields

Using Laplace transforms rules, Eq. (1) is reduced to

Variables G,(s), j = 1, 2, in Eq. (2) represent the Laplace transforms of the yet unknown temperature distributions @,(<), j = 1, 2. A complex independent variable denoted as s replaces the original independent variable <. Knowing the inlet boundary condition at = 0 [i.e., O1(0) = 01, Eq. (2) is solved with respect to the Laplace transform GI ( s j as follows:

<

Ql(S)

+ NTUl

NTU I

Q2 ( 3 )

(3)

In Eq. (3), an explicit form of G2(s) still has to be determined. This can be done involving the other differential equation of the mathematical model [i.e., Eq. (1 1.12)) The same procedure as the one just outlined is applied. Hence,

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

743

and Rl[sOl(s) - 01(0)] i 2 [ s 0 2 ( s ) 02(0)]= 0 Solving Eq. (5) for &(s) and utilizing Eq. (3) for @(s), we get
= s 2 + s N T U l ( l + i 2 R 1 ) Q2(0)
s

(5)

O2(s)

+ NTUj

Substitute Eq. (6) into Eq. (3) to get (s) explicitly in terms of s. Now, applying inverse Laplace transforms on Eqs. (3) and (6), we get (for R l # 1 if i2 = 1)

o1

and

Parameter 02(0) in Eqs. (7) and (8) depends on both design parameters (NTU1, R I ) and the value of i2. The value of 0 2 ( 0 )can be determined for the parallelflow arrangement (i2 = +1) directly from the boundary condition at the fluid 2 inlet [i.e., 02(0) = 0 2 , i = 11. For the counterflow arrangement (for i2 = -l), the value of 02(0) can be obtained by collocating Eq. (8) at the fluid 2 inlet (i.e., at E = l), and solving for 0 2 (0). Consequently,

By inspection of Eq. (9), a generalized algebraic expression for the parameter O2(0) can be formulated to extend the validity of that equation to include parallelflow as follows:

for i2 = +1 (parallelflow) for i2 = -1 (counterflow)


(10)

In Eq. (lo), one should first define the fluid stream direction parameter (i.e., i2 = f l ) , and then select the numerical values for design parameters. Finally, combining Eqs. (7), (8), and (lo), the general solution for temperature distributions for both parallelflow and counterflow exchangers can be written as follows:

744

THERMODYNAMIC MODELING AND ANALYSIS

Inserting i2 = +1 for parallelflow and following temperature distributions:


Flow Arrangement Parallelflow

i2 =

-1 for counterflow into Eq. (1 l), we get the

Flow Indicator i2

Ql(5)

02K)

Counterflow

-1

The temperature distributions for both parallelflow and counterflow arrangements and for several sets of parameters are presented in Fig. E l 1.1B. All the results presented so far for i2 = -1 require that 0 5 R1 < 1 (i.e., R I # 1). In the case of a balanced counterflow heat exchanger, R I = 1 and i2 = -1, the original mathematical model given by the set of equations (1 1.10) and (1 1.12) transforms into

FIGURE El 1.1B Temperature distributions in (a) counterflow and (b) parallelflow heat exchangers.

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

745

with the boundary conditions as given by Eqs. (11.13) and (11.14) for counterflow. The solution of this problem leads to linear dimensionless temperature distributions. The temperature distributions for fluids 1 and 2 are

Discussion and Comments: In this example, it has been shown how to find temperature distributions in a heat exchanger with parallel streams (in geometrical sense). This example demonstrates that both parallelflow and counterflow arrangements represent the two subproblems of a single heat transfer problem that differ only in the stream direction. Still, the character of temperature difference distributions in these two situations is radically different, as shown in Fig. 11.2. Note that this modeling is based on energy balances performed on two control volumes selected arbitrarily from a total of three balances (either for one or the other fluid, or for both fluids). All results of this modeling are the consequence of conservation principles.

11.2.2 True Meaning of the Heat Exchanger Effectiveness


In Section 3.3.1, heat exchanger effectiveness and the maximum possible heat transfer rate qmaxare introduced by definition [see Eqs. (3.37) and (3.42)]. This has provided the basis for the formulation of heat exchanger effectiveness in terms of terminal temperatures of the fluids and their heat capacity rates as in Eq. (3.44). However, that approach requires a priori definition of a hypothetical infinite surface area of the heat exchanger. On the other hand, by knowing the temperature distributions of a given heat exchanger, we can devise the concept of heat exchanger effectiveness without invoking the concept of a hypothetical counterflow heat exchanger of infinite surface. We can show that the definition of heat exchanger effectiveness is obtained using the first law of thermodynamics only (SekuliC, 2000), without invoking explicitly the second law of thermodynamics. The true meaning of heat exchanger effectiveness as a dimensionless outlet temperature of the fluid stream having the smaller heat capacity rate is a direct consequence of this interpretation. Moreover, the maximum possible heat transfer rate
TABLE 11.1 Interpretations of the Meaning of Heat Exchanger Effectiveness
Traditional meaning
E

=4=

Based on a comparison of Defined utilizing the first law of thermodynamics the actual heat transfer explicitly, and the rate exchanged in the - cc(Tc~o - Tc3i) heat exchanger to that second law of Cmin(Th~i - Tcsi) thermodynamics exchanged in an ideal, hypothetical heat implicitly exchanger having U A + 03
Cmin(Th,i - Tc,i)

ch(Th~ i Th'O)

True meaning

E=-

TI.o - Tt ,i
T2.i - T1,i

C, = C,,,

Dimensionless outlet temperature of a fluid with smaller heat capacity rate, C, < C2

The first law of thermodynamics is used explicitly

746

THERMODYNAMIC MODELING AND ANALYSIS

exchanged in a heat exchanger can subsequently be derived, not postulated. The two interpretations are summarized in Table 1 1.1. In Example 1 1.2, these important thermodynamic consequences of the analysis of temperature distributions in a heat exchanger will be illustrated using both parallelflow and counterflow heat exchangers as an example. It should be reiterated that the interpretation given is universally valid, regardless of the complexity of the flow arrangement involved.
Example 11.2 Show that heat exchanger effectiveness and/or temperature effectiveness represent the nondimensional outlet temperature of one of the two fluid streams of a heat exchanger. Use the parallelflow and counterflow arrangements as examples. Demonstrate that heat exchanger effectiveness can be interpreted as a ratio of actual heat transfer rate to the heat transfer rate of a hypothetical exchanger with an infinitely large thermal size (NTU + co), as emphasized in Chapter 3, but without invoking explicitly the second law of thermodynamics.

SOLUTION
Problem Data and Schematic: As presented in Example 1 1.1 Determine: Demonstrate the equivalence between the heat exchanger effectiveness definition and the dimensionless outlet temperature of one of the fluids in a heat exchanger. Assumptions: As invoked in Example 1 1.1. Analysis: The outlet temperature of fluid 1 for parallelflow and counterflow can be obtained from Eq. (1 1) of Example 11.1 at = 1 (the outlet of fluid 1).

<

Substituting iz = + I for parallelflow, we obtain

Substituting i2 = - 1 for counterflow, we obtain

Equations (2) and (3) are identical to Eqs. (1.2.1) and (1.1.1) of Table 3.6. So the dimensionless outlet temperatures are equal to parallelflow (~,,f)and counterflow ( E , ~ ) heat exchanger effectivenesses, respectively. We have obtained expressions for heat exchanger effectiveness without invoking the concept of an ideal heat exchanger. Hence, the true meaning of the effectiveness is simply the dimensionless outlet temperature of the fluid with the smaller heat capacity rate (note that R I= C* in this case). This conclusion is general and valid for any flow arrangement.

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

747

To devise a traditional definition of heat exchanger effectiveness, let us first determine the outlet temperature of a counterflow heat exchanger in the limit of an infinitely large heat exchanger with thermal size NTUl + 00:

Invoking the definition of the dimensionless temperature from Eq. (1 1.9), the following result can be obtained from Eq. (4):

As indicated above, R I = C* 5 1, CI 5 C2 by definition (i.e., fluid 1 has the smaller In that case, from Eq. (3.103), NTUl = NTU. heat capacity rate, C1 = Cmin). Now the heat transfer rate in a countercurrent heat exchanger of NTUl -+ 00 is given as follows using Eq. (5):

The actual heat transfer rate in a two-fluid single-phase heat exchanger of any flow arrangement is as follows:

Finally, dividing the right-hand side of Eq. (7) with qmaxfrom Eq. (6), and comparing the result with the definition of the outlet temperature of fluid 1, one can obtain

This constitutes the proof required. Note that for a counterflow arrangement with R, = C* = 1, the heat exchanger effectiveness becomes E = 8 1(1) = NTU/( 1 NTU); see Example 1 1.1 for the corresponding temperature distributions.

Discussion and Comments: The heat exchanger/temperature effectiveness has its true meaning as a dimensionless outlet temperature of the fluid with the smaller heat capacity rate (heat exchanger effectiveness, E ) or the dimensionless outlet temperature of a given fluid (temperature effectiveness, say P I for fluid 1). The numerical value of the heat exchanger effectiveness is between 0 and 1 and indicates how close the outlet temperature of one fluid can approach the inlet temperature of the other fluid. The traditional meaning of the heat exchanger effectiveness [although a thermodynamic interpretation based on Eq. (8) is perfectly valid and insightful] involves the concept of a hypothetical infinitely large counterflow heat exchanger. Refer to Sekulib (2000) for a detailed discussion concerning an analysis of this approach; discussion of a concept of the thermodynamic efficiency for a heat exchanger is given in Section 11.6.5.

748

THERMODYNAMIC MODELING A N D ANALYSIS

11.2.3 Temperature Difference Distributions for Parallelflow and Counterflow Exchangers Let us now address the magnitude of the local temperature difference between fluids 1 and 2 ( A T = ITI - Tzl)in either a parallelflow or counterflow exchanger.+ We need this information to better understand the influence of temperature distributions on E , P, or F. The distribution of local temperature differences for parallelflow and counterflow heat exchangers can be determined in a general form by utilizing the corresponding temperature distributions: for example, Eq. (1 1) of Example 11.1. The temperature difference distribution is as follows (see Problem 11.10 for details):

Note that Eq. (11.15) is valid for both parallelflow (i2 = +1) and counterflow (i2 = -1) arrangements for 0 5 R I < 1. In Fig. 11.2, a graphical representation of Eq. (1 1.15) is given for both counterflow and parallelflow arrangements and for several

______-----Parallelflow -

Counterflow

AO

; '

0.0

0.2

0.4

0.6

0.8

1 .o

FIGURE 11.2 T e m p e r a t u r e difference distributions for a c o u n t e r f l o w exchanger and a parallelflow exchanger w i t h R I = 0.6.
The terms para//elfloii3 and counterflow are used throughout the book following the practice by Kays and London (1998), which has been extensively used throughout the world for almost last five decades. As demonstrated in this section using a straightforward first law analysis, a more favorable terminology from a semantic point of view would be unidirectional and bidirecrional flows (i.e., in both arrangements, fluid streams are flowing parallel in a geometric sense but oriented in the same or opposite directions).

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

149

values of NTUl and R I = 0.6. Refer to SekuliC (2000) for more detailed data. A study of these temperature difference distributions reveals the following conclusions.
0

As we have learned in Chapter 3, counterflow is the best and parallelflow the worst flow arrangement from the effectiveness viewpoint for given NTU and C* (or NTUl and R,). For the same heat capacity rate ratio (say, R I = 0.6), and in particular at large NTUl values, the magnitude of temperature difference change along the flow direction (0 is substantially larger for parallelflow than it is for counterflow (compare corresponding curves in Fig. 1 1.2). The two temperature distributions are identical for R1 = 0 (not shown in Fig. 11.2) and differ more at large heat capacity rate ratios. For a given NTU for counterflow, A@ features a more pronounced variation along [ as R I (or C*) decreases from 1 to 0. This means that there will be larger local temperature differences with decreasing R I (or C*) if all other parameters remain the same. If the temperature difference distributions are close to each other or identical, they will have similar or identical effectiveness. If the temperature difference distributions differ substantially, the heat exchanger effectiveness will differ considerably as well.

In the limiting case of R I = 1, the temperature difference distribution for counterflow arrangement is uniform throughout the exchanger, A@ = Q2 - Q1 = I / ( 1 + NTUI), as derived in Example 11.1. For the same heat capacity rate ratio in the parallelflow arrangement, the change in local temperature difference is the largest possible. Comparing the corresponding heat exchanger effectiveness (see Figs. 3.7 and 3.8) for these conditions, one can easily conclude that the temperature effectiveness of these two flow arrangements differs the most. The distribution of local temperature differences has a profound influence on exchanger effectiveness (either E or P I ) .A finite temperature difference between the two fluid streams is the driving potential for heat transfer, but large temperature differences lower the exchanger or temperature effectiveness, and ultimately may contribute to a lower system efficiency, as we demonstrate in subsequent sections.

11.2.4 Temperature Distributions in Crossflow Exchangers

A model of a crossflow arrangement with/without mixing provides a good example of how a simple geometric configuration of two fluid streams in thermal contact may lead to two-dimensional temperature fields with a very complex relationship between the design parameters. In addition, a study of temperature distributions within a crossflow heat exchanger illustrates how mixing may influence an outcome of the heat transfer process. This insight is very important for an assessment of the influence of mixing on a reduction in exchanger thermodynamic performance, as demonstrated in Section 11.5. In Fig. 1 1.3, a schematic of the main features of the geometry and control volumes in a crossflow heat exchanger is presented. Fluids 1 and 2 flow perpendicular to each other over a heat transfer surface that separates them. This flow arrangement is discussed in Section 1.6.1.3 (see Fig. 1.53). All the assumptions of Section 3.2.1 are invoked here. Due to heat transfer across the heat transfer surface, both fluids will change their temperatures in either one or both flow directions, depending on the presence or absence of mixing, respectively.

750

THERMODYNAMIC MODELING AND ANALYSIS

Four distinct situations are possible with respect to mixing of the fluids, as emphasized in Section 1.6.1.3. These are also shown graphically in sketches for Eqs. (II.lHII.4) of Table 3.6. In Table 11.2, various temperature distributions are shown to be either oneor two-dimensional [Oj = f ( x or [) or 0, = f ( x ,<), j = 1,2], depending on fluid mixing. Our goal now is to show how one can formulate the models and subsequently solve them to find these temperature fields and/or corresponding outlet temperatures. As a byproduct of this analysis, one can easily devise temperature and/or heat exchanger effectiveness. In Tables 3.3 and 3.6, the formulas for the heat exchanger effectiveness are listed. Here we discuss the analytical models for determining both temperature fields and effectiveness. This will provide an insight into the influence of fluid mixing on heat exchanger effectiveness, discussed in Section 11.3. Referring to Fig. 11.3, one can write energy balances for control volumes as follows:

fluid enthalpy rate into the control volume

-.
dritlc,,lTI

-drit,cp,l(T,+gdx) 0

v dq
heat transfer rate from fluid to wall

=O

(11.16)

fluid enthalpy rate out of the control volume

'

and
dh2~p,2T2
fluidinto the control volume

v
heat transfer rate in from lo fluid

dq

-dljl2cp,2
fluid enthalpy rate out of the control volume

(11.17)

Note that drhj cp,i = dCj, j = 1, 2, in Eqs. (1 1.16) and (1 1.17) implies that constant thermophysical properties assumption is invoked. It is assumed that no mixing takes place on either side of the heat transfer surface. Consequently, both fluids will have twodimensional temperature fields. In Eqs. (1 1.16) and (1 1.17), dq represents heat transfer by convection from the hot fluid to the wall; and in the steady-state formulation, that heat will be transferred by conduction through the wall and by convection to the cold fluid.

Heat transfer surface, T ,

FIGURE 11.3 Energy balance control volumes for crossflow arrangements.

TABLE 11.2 Models of Crossflow Arrangements'


~~~

Model Inform ation Temperature field Unmixed-Unmixed C,,, Fluid Unmixed 01 = 4(x) C,,, Fluid Unmixed Mixed-Mixed

4 = &(x,C)
02 = 02(x,C)

e2 = e2(x,c)

=4 ( x , l )

4 = &(x)
02 = M

0 2 = &(C)

C)

Independent variables Boundary conditions Schematic

NTU 05x I 0I c 5 C* .NTU 0 1 (0,C) = 1 02(x,0) = 0

0I x 5 NTU o5cI C* .NTU O,(O) = 1

e2(x,0) = 0
Fluid
2

Fluid

Fluid 1 d2

r l
Fluid 2

O,(O) = 1 O,(O) = 0

F l d q T k u i d

"The dimensionless variables are defined as follows: 0, = (T, - T 2 , , ) / ( T ,~ ,Tz,,), x = (x/Ll)NTU, and C = ( y / L , ) C * .NTU. Horizontal and/or vertical lines in schematic figures symholize flow of the respective fluid through the flow areas characterized with one-dimensionality (along the flow direction). The absence of lines in the direction of flow symbolizes fluid mixing in the direction transverse to the flow direction. Fluid 1 is assumed to have a smaller heat capacity rate C,,,,,.

752

THERMODYNAMIC MODELING AND ANALYSIS

Using the rate equations for convection and conduction, dq from the hot fluid to the cold fluid can be expressed as follows:
d q = rlo.lhl(T1 - Tl,.,,)dXdY = k , .
convection from hot fluid to wall

TW,l - TW,2
convection from wall to cod fluid
*

"

'conduction within * ' the wall

(11.18) Equation (1 I . 18) indicates that there is neither energy generation nor axial conduction through the separating wall, as idealized in Section 3.2.1. The products h, qo,,AT,, j = 1, 2 of Eq. (1 1.18) represent the heat transfer rates exchanged per unit heat transfer area between fluids 1 or 2 and the wall separating the fluids. In the case where the thermal resistance of the separating wall is neglected, only convective terms exist in Eq. (1 1.18), = T,l,,2 = TIC (Le., the heat transfer surface has a uniform wall temperature with orthogonal to the flow directions). Note that the balances presented by Eqs. (1 1.16)(1 1.18) do not involve the overall heat transfer coefficient. They are formed by applying the thermodynamic convention for each of the thermal energy flow rates (positive if entering the system, otherwise negative). Note also that the assumptions of uniform distribution of the heat transfer area and uniform wall thermal resistance are invoked. Equation (1 1.18) is rewritten as follows:

(11.19) Adding up the temperature differences from three equations of Eq. (1 1.19), defining dA = dxdy, and using Eq. (3.18) for definition of the overall heat transfer coefficient U , but neglecting the fouling thermal resistances [since we did not include them in the formulation of Eq. (1 1.18), which we could have included readily if desired], we get
dq = UdA(T1 - T2)

(11.20)

Substituting Eq. (1 1.20) into Eqs. (1 1.16) and (1 1.17) and simplifying, we can get the following partial differential equations: (1 1.21)
, i T z , ~ ) , with j = 1, 2, x = (x/LI)NTU, and where ej = (q - T 2 , i ) / ( T lC = ( y / L 2 ) C * NTU. . The number of heat transfer units NTU is based on the heat transfer surface between the fluids defined (for the sake of clarity) as the product of Ll and L2 (see Fig. 11.3). Note that UA is distributed uniformly throughout the exchanger due to assumptions of uniformity of heat transfer surface and thermal resistances. The same holds for the heat capacity rates. Hence, NTU is based on any (hot or cold fluid side) heat transfer surface on which U is defined. The definition of dimensionless temperatures 0, is complementary to the definition of the dimensionless temperature 0 of Eq. (1 1.9). This flexibility in defining dimensionless temperature allows an analyst to define the heat exchanger effectiveness as either a dimensionless outlet temperature of the fluid

MODELING A HEAT EXCHANGER ON THE FIRST LAW OF THERMODYNAMICS

753

with smaller heat capacity rate or as its complementary value. Note also that both dimensionless temperatures 0, are assumed to be locally dependent on both independent coordinates x and C:
01 = 61 ( X l O

02 = 02(x1 C)

(1 1.22)

where the ranges of independent variables take the values


0 5 x I NTU 05

C I C*.NTU

(1 1.23)

Two boundary conditions (for uniform inlet temperatures) accompany the set of Eqs. (1 1.21):

~I(ol0= 1

0 2 ( X > O )= 0

(1 1.24)

The set of equations Eqs. (1 1.21) and (1 1.24) represents the mathematical model of a crossflow heat exchanger. Four particular cases of crossflow (see Section 1.6.1.3, and Tables 3.6 and 1 1.2) differ from each other with respect to the presence or absence of fluid mixing on each fluid side within the heat exchanger core (see Problems 11.5 and 1 1.6). In Table 1 1.2, a summary containing all four models, deduced from Eqs. (1 1.21)411.24), is presented. Each of these models can be solved and closed-form analytical solutions can be obtained using various solution methods (see Section 3.11). The solution of the general unmixed-unmixed case is asked in Problem 11.2. A particular case of an unmixed-mixed crossflow arrangement is considered in detail in the following example. The mixed-mixed case is considered in Problem 11.7. Example 22.3 Determine temperature difference fields in a heat exchanger with a mixed-unmixed crossflow arrangement. Assume that the fluid with the smaller heat capacity rate is mixed. SOLUTION
Problem Data and Schematic: The flow arrangement under consideration corresponds to the model and schematic given in the third column of Table 11.2 (C,,, fluid unmixed). Determine: Temperature difference as a function of both axial and transverse coordinates

(x and C, as defined in Table 11.2).

Assumptions: The assumptions are as presented in Section 3.2.1. Analysis: We first determine the temperature fields for both fluids followed by a temperature difference distribution relationship within the heat exchanger core. The analytical model consists of two equations, one partial and one ordinary differential equation, as presented in the third column of Table 11.2 (the details of the model development are the subject of Problem 11.6). Uniform temperatures are considered at inlets as corresponding boundary conditions. The solution to this analytical model will provide the desired temperature fields. Let us first solve the partial differential equation for fluid 2. Subsequently, using the temperature field solution for fluid 2 and replacing its explicit form in the ordinary differential equation for fluid 1, we will find the temperature

754

THERMODYNAMIC MODELING AND ANALYSIS

distribution for fluid I . Finally, the difference between the two fluid temperatures will provide the solution of the problem. The solution of the partial differential equation for fluid 2 from Table 11.2 can be obtained as follows using the Laplace transforms technique:

where s represents a complex variable that replaces O2(x,O) = 0 from Table 2, we get

C. Rearranging

Eq. (2) with

An inverse Laplace transform of Eq. (3) provides the temperature field for fluid 2:

Note that the explicit form of 6$ (x) still has to be determined. The ordinary differential equation for fluid 1 of Table 11.2 can now be written as follows:

After determining the integral on the right-hand side of Eq. (5) and rearranging, the differential equation for the fluid 1 temperature distribution becomes

where X = [l - exp(-C*

. N T U ) ] / ( C * . NTU). The boundary condition for Eq. (6) is


O,(O) = 1
(7)

The solution of a simple problem defined by Eqs. (6) and (7) is

01 (x) = e-X

(8)

So the temperature field of fluid 2 can be obtained by introducing the temperature distribution of fluid 1 given by Eq. (8) into Eq. (4):

e2(x,c) = (1 - e-C)e-xX

(9)

Finally, the relationship for the temperature difference distribution can easily be determined from Eqs. (8) and (9) as

W x ,C) = 01 (x) - 02(x, C) = e- (X+C)

(10)

IRREVERSIBILITIES IN HEAT EXCHANGERS

755

Discussion and Comments: As expected, the temperature distribution for unmixed fluid 2 is two-dimensional and that for mixed fluid 1 is one-dimensional, both dependent on X , which in turn depends on NTU and C*. Knowing the temperature distribution of the fluid with the smaller heat capacity rate (fluid 1 in this case), one can easily determine

heat exchanger effectiveness (solve Problem 11.3 for understanding the details). Similarly, the analysis of a crossflow arrangement for the mixed fluid having a larger heat capacity rate can be performed, and heat exchanger effectiveness can subsequently be determined (see Problem 11.4). Even more complex situations with nonuniform inlet temperatures are the subject of Problems 1 1.8 and 1 1.9.

11.3 IRREVERSIBILITIES IN HEAT EXCHANGERS

Important phenomena that shape the heat transfer and flow characteristics within a heat exchanger are (1) heat transfer at finite temperature differences, (2) mixing and/or splitting of the fluid streams, and (3) fluid flow friction phenomena; additional phenomena when present are phase change, flow throttling, and so on. The first two phenomena influence temperature distributions, and the third the flow friction characteristics on each fluid side of a heat exchanger. Thermodynamics teaches us (Bejan, 1988) that these processes are accompanied by entropy generation, an indicator of undesirable thermodynamic irreversibilities that diminish the thermal performance. So thermodynamic irreversibility is an inevitable by-product of these processes and a principal cause of exchanger/system performance deterioration. Some of the irreversibilities associated with heat transfer and fluid flow are (Gregorig, 1965; Sontag and Van Wylen, 1982):
0

0
0

0
0

Heat transfer across a finite temperature difference (including both heat transfer between the fluids and heat transfer across the heat exchanger boundary, i.e., heat leak and/or gain to/from surroundings) Mixing of dissimilar fluids (dissimilar with respect to p , T, and/or composition) Fluid friction and flow impact Phase change where initial conditions are not in equilibrium Flow throttling

In this section we focus on identification and quantitative evaluation of the three dominant irreversibilities in a heat exchanger: (1) irreversibility caused by a finite temperature difference, (2) irreversibility caused by fluid mixing, and (3) irreversibility caused by fluid friction. We evaluate these irreversibilities in terms of entropy generation. This analysis will assist us in assessing the quality of heat transfer and associated phenomena in heat exchangers that cannot be evaluated and explained by the analysis presented in Sections 3.3 through 3.8. This analysis requires simultaneous use of both the first and second laws of thermodynamics and introduction of the concept of exergy (see Section 1 1.6.4). Let us first review very briefly the concepts of irreversibility, entropy, entropy generation, and exergy before we start the foundation of thermodynamic analysis of heat exchangers. For further details on these concepts and related thermodynamics background, refer to Bejan (1988) and Moran and Shapiro (1995). Thermodynamic irreversibility (simply referred to as irreversibility) is a term used to describe a natural tendency of any real system not to be able to revisit the same sequence

756

THERMODYNAMIC MODELING AND ANALYSIS

of states during a reverse change of state from the final to the initial state without additional energy interaction(s). An additional energy interaction is due to an absence of reversibility in a real world when thermal phenomena are involved. In practical terms, this means that the presence of irreversibilities is accompanied by thermodynamic losses, ultimately leading to poorer thermal performance than predicted by an idealized reversible process. In a narrower sense, the same term is used to describe the losses in energy terms iirr caused by the presence of irreversibilities. The value of irreversibility cannot be negative and it is nor a system property as discussed next. Irreversibility can be expressed in energy terms as a product of entropy generation and a temperature weighting factor To (i.e., Zirr = To$irr).It can be shown that in many engineering applications the weighting factor can be interpreted as the temperature of the surroundings, which is identified as a thermodynamic reference state for measuring the thermal energy potential of the system at hand. It should be noted that entropy generation is not a property of a system, while the entropy S is. Entropy is defined as a system property by a statement that its change in an ideal, reversible process must be equal to the transfer of an entity J d q / T that accompanies any heat transfer dq across the system boundary where the local temperature is T. Hence, this abstract system property indicates that heat transfer must be accompanied by an entropy change. As a consequence, a reversible adiabatic process can be identified by zero entropy change. If a process is not reversible (as with any heat transfer across a finite temperature difference), the situation is radically different. Entropy change AS is either equal (reversible process) or larger (irreversible process) than the entropy transfer dq/ T , nonproperty)+ that accompanies heat transfer dq, the difference being attributed to entropy generation [see Eq. (1 1.36)]. The amount of entropy generation is the quantitative measure of the quality level of energy transfer. Entropy generation of zero corresponds to the highest quality of energy transfer and/or energy conversion (a reversible process), and entropy generation greater than zero represents poorer quality. All real processes are characterized by entropy generation greater than zero. The concept of exergy or available energy E is introduced to describe the maximum available energy that can be obtained from a system in a given state. Each fluid stream that enters or leaves a heat exchanger carries exergy rate. Due to irreversible processes in a system (e.g., a heat exchanger), the available energy of a fluid decreases and the difference between the input and output exergy rates is equal to the lost exergy (lost available energy), which in turn is identical to the irreversibility in energy terms (a GuyStodola theorem, i.e., exergy destroyed = lost available energy = temperature weighting = To.$rr). factor x entropy generation; that is, A& = klost

sir,

(s

sir,

11.3.1 Entropy Generation Caused by Finite Temperature Differences

Temperature and temperature difference distributions within the heat exchanger influence thermodynamic irreversibility. Thermodynamics teaches us that a measure of the efficiency of any thermal process can be assessed by gaining an insight into the irreversibility level incurred in an associated heat transfer process (BoSnjakoviC, 1965). This irreversibility can be identified by determining the corresponding entropy generation. So there is a deeper rationale for turning our attention toward temperature differences within a heat exchanger to determine heat transfer performance behavior. The driving potential for heat transfer in a heat exchanger is the finite local temperature difference
'Note that the entropy is a system property and entropy transfer or entropy generation is no1 a system property.

IRREVERSIBILITIES IN HEAT EXCHANGERS

757

between the fluids exchanging heat, and we should expect that it greatly influences the exchanger effectiveness as well. Consequently, it is plausible to conclude that these temperature differences are related to both the exchanger effectiveness and thermodynamic efficiency of an exchanger (one such thermodynamic figure of merit is defined in Section 1 1.6.5). How these temperature differences influence the irreversibility level is explained next in terms of entropy generation. The thermodynamic irreversibility manifested within a heat exchanger as an adiabatic open system can be identified in terms of entropy generation by total entropy change (entropy measure of irreversibility, i.e., entropy generation rate $,) of both fluid streams: Sir, = A S = r i r l As1

+ rir2 As2

(1 1.25)

Now we will evaluate Sir,only due to finite temperature differences, considering the fluids as pure simple single-phase compressible substances. Since ds = dh/T (where h is the specific enthalpy), the entropy rate change for fluid 1 in a heat exchanger operating under steady-state conditions for an ideal gas or an incompressible liquid is given by

Similarly, the entropy rate change for fluid 2 in the exchanger will be

~)T2,o ~ m2 As2 = ( r i r ~In T2.i

(11.27)

Note that we do not need to distinguish at this point whether the fluid is hot or cold. Moreover, this distinction is not necessarily relevant for calculating the entropy measure of irreversibility. What matters, though, is that the two fluids have dzferent temperatures. Hence, the concepts of hot and/or cold will not necessarily be used here. Consequently, Eq. (1 1.25) can be rewritten as follows: (1 1.28) The two terms in Eq. (1 1.28) have the opposite signs since the fluids have different temperatures # T2,i, i.e., either 5 and T2,0 2 T2,i or 2 TI,^ and T2,0 5 T2,i).Two important thermodynamic points have to be reiterated here. First, the fact that the two fluids have different temperatures is of far more importance than that one may conveniently be described as hot and the other as cold. This is because the hot/cold dichotomy is introduced by convention. In a heat exchanger, as will be demonstrated later (see Section 11.4.3), the same fluid may change the role of a hot/cold fluid side over some flow length! So, in this chapter, we will, as a rule, refer to a fluid as fluid 1 or fluid 2 whenever a general case has to be considered in which any of the two fluids may either be hot or cold. If a fluid is identified as having higher/lower temperature (such as in a particular given example), we denote as the hot fluid that has a temperature at its inlet porr higher than the temperature of other fluid at its inlet port. Second, a more advanced thermodynamic analysis advocated in this chapter involves the concept of entropy; hence

758

THERMODYNAMIC MODELING AND ANALYSIS

it involves not only temperature differences but also temperature ratios and the products of absolute temperature and entropy differences [see, e.g., Eq. (1 1.28) or (1 1.53)]. As a consequence, proper care must be taken regarding the use of absolute temperatures (K or O R ) for all temperatures associated with entropy and also exergy later. To emphasize this fact, we use in this chapter, as a rule, temperatures on the absolute Kelvin (or Rankine) scale and not on the commonly used degree Celsius (or Fahrenheit) scale. The heat transfer rate between the two fluid streams in thermal contact under adiabatic conditions is equal to the respective enthalpy rate changes (see Chapters 2 and 3):

For better clarity, we consider fluid 1 as the hot fluid and fluid 2 as the cold fluid. Hence, - Th,o), the enthalpy rate changes for the hot and cold fluids are Ahh = cp,h(Th,i Ahh = cp,h( Tc,o- Tc,i)for Eq. (1 1.29). Changing the subscripts 1 and 2 of Eqs. (1 1.28) and (1 1.29) to h and c, combining them, and rearranging, we get,
(1 1.30)

where (11.31) represents the log-mean temperature of the hot fluid as defined using inlet and Here ThJm outlet temperatures Th,j and Th,o. The Tc,lm is defined similarly. In contrast, the arithmetic = (Th,i Th,,)/2 and mean temperatures of the hot and cold fluids are Th,m Tc,m = ( Tc,i+ Tc,o)/2, and the log-mean temperature diference between hot and cold fluids in a heat exchanger is given by Eq. (3.172). The entropy generation is related to the difference in the fluid temperatures. Equation (1 1.30) is written for the exchanger as a whole. On the local level, entropy generation is related to local temperature differences [such as Eq. (1 1.15)]. Hence, the diference between mean temperatures of two fluids [the numerator of Eq. (1 1.30)] directly influences the entropy measure of the irreversibility manifested within the heat exchanger. As a consequence, the heat exchanger irreversibility for a given heat transfer rate can be reduced by reducing temperature differences between the fluids, which in turn will increase the exchanger effectiveness E.+ A heat exchanger characterized by smaller temperature differences between the fluids generates a smaller irreversibility in a given system compared to a heat exchanger (for the same heat transfer rate) that has larger temperature differences between the fluids. Since the entropy measure of irreversibility is related directly to thermodynamic system efficiency (see Section 11.6.5), this statement leads to an anticipated conclusion about the possible detrimental influence of this source of irreversibility on the overall system efficiency. Thermodynamic irreversibility represented by entropy generation as in Eq. (1 1.28) can be formulated in terms of heat exchanger thermal design parameters. Using the

+Note that the log-mean temperature difference AT,, of Eq. (3.172) is proportional to (Th,,, - Tc,lm). Hence, a smaller value of AT,, means a larger value of E for an exchanger.

IRREVERSIBILITIES IN HEAT EXCHANGERS

759

definitions of heat exchanger effectiveness and heat capacity rate ratio Eqs. (3.44) and one can show that (3.56)] and considering CI = Cminr
-= "" 1 +~(19-' Tl,i

1) = 1 - P 1 ( K 1 - 1)
T2,i

=1

+ C * E (~ 1) = 1 + RIP1(19 - 1)
(11.32)

where 29 = Tl,i/Tz,i represents the inlet temperature ratio. Substituting these expressions in Eq. (1 1.28) results in

Sir, = S*
Cmax

= C* ln[l

+ ~ ( 8 -l I)] + l n [ l + C*E(O - I)]

(1 1.33a)

where # 0 for 29 # 1 and $,,= 0 for 29 = 1. Note that normalizing Sirrby Cmax or C2, as indicated in the leftmost term of Eq. (1 1.33), is a matter of arbitrary choice. The entropy generation is equal to zero for the inlet temperature ratio equal to 1. Entropy generation [Eq. (1 1.33)] for different flow arrangements is different for the same inlet temperature ratio, heat capacity rate ratio C* or R1, and NTU (SekuliC, 1990b). This is because of different heat exchanger effectivenesses, E = E (NTU, C*), for different flow arrangements [and fixed values of NTU and C* (or NTUl and RI)]. It should be emphasized that the control volume for the entropy generation of Eq. (1 1.33) [see Eq. (1 1.25)]is drawn outside the exchanger core or matrix through inlet/outlet ports. Hence, the S* expression of Eq. (1 1.33) is valid for an exchanger with anyflow arrangement by employing its appropriate E-NTU or P-NTU formula. Some additional features of Eq. (1 1.33) are discussed in Section 1 1.4.1.

sir,

11.3.2 Entropy Generation Associated with Fluid Mixing


Fluid mixing in a heat exchanger causes thermodynamic irreversibility and generates entropy, leading to a reduction in the thermodynamic efficiency of the heat transfer process, thus reducing the heat exchanger effectiveness. In general, the mixing of fluids that are dissimilar with respect to their composition and/or state variables is an irreversible process. These dissimilarities may be mechanical (pressure gradients), thermal (temperature gradients), and/or chemical (chemical potentials). Entropy generation associated with mixing depends on the degree of dissimilarity between mixing fluids. The irreversibility associated with a mixing process is due to (1) a process of intermingling molecules of different substances, (2) energy interchange between the same or different substances or within the mixing substances, (3) heat transfer between surroundings and mixing substances, and (4) viscous dissipation effects. For the analysis of many heat exchangers, thermal dissimilarities (due to temperature gradients on each fluid side in the transverse direction) of two or more fluid streams during mixing are of primary interest. For example, in a crossflow heat exchanger, heat transfer between two fluids causes the presence of local temperature nonuniformities in any given flow cross section. However, a fluid flowing through a nonpartitioned flow passage (i.e., a mixed fluid stream side) is characterized by an important feature. The unrestrained mixing attenuates temperature nonuniformities at a given cross section of

760

THERMODYNAMIC MODELING AND ANALYSIS

, I

A virtual stream

..

. .. .

. .. . .

temperature inlet, Ma

Uniform temperature

partitioned manifold), fluid unmxed FIGURE 11.4 Flow passages with fluid mixing: (a) passage or duct with fluid mixing; ( b ) outlet manifold; (c) two adjacent flow passages with/without fluid mixing.

the mixed fluid and the available thermal potential is destroyed. This is certainly an irreversible phenomenon that leads to a corresponding entropy increase. A practical consequence of this thermodynamically detrimental process is the equalization of fluid local temperatures across the flow passage cross section, ultimately leading to a reduced heating/cooling manifested at the fluid exits. As a consequence, the respective temperature effectiveness (or the heat exchanger effectiveness) is also reduced. Let us consider a fluid stream being mixed while flowing right to left through a duct, as shown in Fig. 1 1 . 4 ~The . objective of the analysis is to determine entropy generation associated with fluid mixing using a very simplified approach. This situation is often present in heat exchangers. For example, in the outlet header, mixing is accomplished between different streams of the same fluid ideally with no heat transfer with the surroundings as shown in Fig. 11.4b. In the heat exchanger core region, a mixed fluid exchanges heat with the other fluid, either mixed or unmixed. A significant simultaneous for the unmixed-mixed case) prevent us from heat transfer and mixing (see Fig. 1 1 . 4 ~ determining the sole contribution of mixing to the total irreversibility for a control volume under consideration. That is the reason why we formulate a more general case but consider a mixing-only case as follows. A mixed fluid (fluid 1 in Fig. 1 1 . 4 ~flows ) through the passage while simultaneously being mixed in a direction transverse to the overall flow direction. For the sake of clarity (but with an inevitable loss of rigor), let us assume that resultant heat exchange between the fluid and the environment can be modeled as heat transfer between each of n virtual streams of the mixed fluid at the inlet that merge into one mixed stream at the outlet. Mass rate, energy/enthalpy rate, and entropy rate balances (continuity, energy, and entropy equations) for the control volume of the bulk flow of fluid 1 are as follows:

IRREVERSIBILITIES IN HEAT EXCHANGERS

761

Continuity equation:

Energy equation:

Entropy equation: where qj, j = 1, n, represent the equivalent heat transfer rates between virtual streams (having average individual bulk temperatures T/ along the respective flow paths) and the surroundings (the other fluid side). Note that Sir, > 0 in Eq. (1 1.36) is a consequence of the second law of thermodynamics for a real system. For a steady-state flow, Eqs. (1 1.34H11.36) reduce tot

(1 1.37)
moh0 =

c
n

mj,ihj,i+

c
n

qj

(11.38) (11.39)

Taking into account the change of local temperatures along the flow paths would, indeed, require writing the balances in a differential form and integrating them along the flow path. The form of these relationships would depend on actual heat transfer conditions. For isolating the mixing effect only, we consider an adiabatic mixing case. The simplest physical situation of practical interest corresponds to the conditions encountered in the headers/manifolds or in the parts of a heat exchanger where the heat transfer q can be neglected (as is truly for example, in the exit zone of a TEMA J heat exchanger; Fig. 1 1.6). In such an adiabatic situation, Eqs. (1 1.38) and (1 1.39) can be simplified as follows:
(1 1.40)
n n

(11.41) For a simple compressible substance, using the expression of Eq. (1 1.26) for mj Asj for each stream, Sir, of Eq. (1 1.41) reduces to
(1 1.42)

' qj is not shown in Fig. 11.4.

762

THERMODYNAMIC MODELING AND ANALYSIS

Hence, = 0 for uniform inlet temperature ( = Ti = To,j = 1, n) and # 0 when thermal dissimilarity is present (i.e., nonuniform temperature q,i# To). For example, mixing the two streams of fluid 1 at the exit of a 1-2 TEMA J shell-and-tube heat exchanger (see Fig. 11.6 where To = TI,o, and q,i are denoted as T;,oand TI(,) is a source of irreversibility due to the fact that these outlet temperatures of the shell fluid (fluid 1) leaving the two zones A and B are not the same. From Eq. (1 1.42), it is obvious that entropy generation caused by mixing would never be equal to zero if thermal dissimilarity is present while mixing the streams, even for a single fluid. The mixing process actually eliminates the presence of local temperature differences; hence it is an inherently irreversible process. This statement holds true regardless of the presence or absence of heat transfer to the environment (or other fluid) during mixing. Consequently, it is expected that a heat exchanger featuring mixing of either fluid within the heat exchanger core or within the header zones should have less effectiveness compared to an exchanger with the same design parameters but without mixing.

sir,

sir,

11.3.3 Entropy Generation Caused by Fluid Friction

The importance of fluid pressure drop and fluid pumping power P in the heat exchanger is discussed in Chapter 6. One of the important components of the fluid pressure drop in the heat exchanger is the fluid friction associated with flow over the heat transfer surface. We derive the irreversibility associated with this fluid friction in this section. Since the control volume is drawn at the inlet and outlet pipes/tanks, this analysis does take into account the contribution of both skin friction and form drag that is important in many exchangers. To identify the irreversibility caused only by fluid friction, let us assume a fluid flowing through a flow passage of an arbitrary cross section. The flow is caused solely by the pressure difference between the two points along the fluid path. The entropy generated with such a flow is equal to the entropy change between the two points along the flow path, say between inlet and outlet. If the enthalpy change contribution to entropy change can be neglected (steady and adiabatic flow), the entropy change is as follows using the T ds relationship: dh = 0 = T ds w dp (where s and w are specific enthalpy and specific volume, respectively):

(1 1.43) For an ideal gas flow, Eq. (1 1.43) reduces to


Pi

(1 1.44) where pressure drop A p = p i - po 2 0. For an incompressible fluid (liquid) flow, under nonadiabatic conditions entropy generation caused by fluid friction is as follows as discussed by Roetzel in London and Shah (1983).
ro

( 1 1.45)

THERMODYNAMIC IRREVERSIBILITY AND TEMPERATURE CROSS PHENOMENA

763

In both Eqs. (11.44) and (1 1.45), we have $ , ,= 0 for Ap = 0 and $,,# 0 for Ap # 0. Hence, the entropy generation caused by fluid friction is never equal to zero for Ap > 0, as in a heat exchanger. In a heat exchanger with two fluids, the irreversibility contribution of each of the two fluids has to be included [i.e., two terms of the form of Eq. (1 1.44) or (1 1.45) have to be calculated].

11.4 THERMODYNAMIC IRREVERSIBILITY AND TEMPERATURE CROSS PHENOMENA


In Section 11.3.1 we have demonstrated that heat transfer, accomplished at finite temperature differences in a heat exchanger, must be accompanied by entropy generation. This entropy generation is a function of heat exchanger design parameters [see Eq. (1 1.33) and Problem 1 1.1 I]. Let us now explore this relationship to relate thermodynamic performance of a heat exchanger to its heat transfer and design parameters.

11.4.1 Maximum Entropy Generation


Let us rewrite Eq. (1 1.33) in a symbolic form as a function of relevant design parameters as follows:
-= S* = f(C*, E , 29) = f(C*, NTU, 29, flow arrangement)
~ m a x
Sir,

(1 1.46~)
(1 1.46b)

-=
c 2

S* = f ( ~ ,P,, , 0) = f ( ~ NTU,, ~ , 8, flow arrangement)

The second equality in Eq. (1 1.46) is written by taking into account Eq. (3.50). So S* is a function of the heat capacity rate ratio, NTU, inlet temperature ratio, and flow arrangement. In Fig. 11.5, this relationship is presented for counterflow and parallelflow arrangements for fluids having equal heat capacity rates (C* = I), and an inlet temperature ratio equal to 0.5 (i.e., 29 = 0.5). It can be shown that corresponding curves for numerous other flow arrangements will be located between the two limiting cases presented in Fig. 11.5 (SekuliC, 1990b). It should be emphasized that these curves (except for parallelflow) have at least one distinct maximum, as explained next. For an exchanger at small NTU values, when NTU + 0, the magnitude of entropy generation will tend to zero (Le., S* + 0). This is certainly an expected result because at NTU = 0 there is no heat transfer since U A = 0 despite the temperature potential for it (represented by the given inlet temperature difference). On the other side, if NTU + m, the temperature differences along the heat exchanger tend to their minimum possible values (e.g., AT = 0 for C* = 1 for a counterflow exchanger). Consequently, S* decreases to a limiting asymptotic value (equal to zero for C* = 1 for a counterflow exchanger). Hence, a curve having minimum values at both ends will have at least one maximum value in between 0 < NTU < co. This analysis thus provides the following conclusions for many flow arrangements having only one maximum for S* [including counterflow but excluding parallelflow (SekuliC, 1990a); see Section 11.4.3 for an exception; refer to Shah and Skiepko (2002) for other exceptions]:

764

THERMODYNAMIC MODELING AND ANALYSIS

FIGURE 11.5 Entropy generation in parallelflow and counterflow exchangers with C* = 1.

> 0 at small NTU


dS*
=0

NTU-0

Iim S* = O (1 1.47)

at NTU = NTU* = 0 at large NTU

S* = S2ax> 0 at N T U = NTU* lim S* = Ski,,, 1 O


NTU-cc

< 0 at NTU* < NTU < m


The most interesting feature of the heat transfer irreversibility behavior implied by Eq. (1 1.47) is an existence of at least one maximum of S* for a finite-size heat exchanger at NTU* (or NTUT). Let us determine the value of NTU* and the corresponding effectiveness at that operating point (the same can be done for the number of transfer units defined as N T U l and the corresponding temperature effectiveness PI). The entropy generation maximum is characterized byt
-=0 dNTU

as*

at NTU
-%ax

=NTU*

(1 1.48)

Performing the calculation as indicated in Eq. (1 1.48) on Eq. (1 1.33), one can show (see Problem 1 1.13) that the maximum of entropy generation corresponds tot:

I t can be shown that @S*/i3NTU2 < 0 at NTU = NTU* (SekuliC, 1990a). Explicit expressions for NTU in terms of E and C* are available only for a limited number of flow arrangements in terms of C*can be as outlined in Table 3.4. For these arrangements, explicit formulas for NTU' (NTU at SZaX) from Eq. (11.49). For the related formulas for countercurrent and crossflow obtained by substituting E at SZOx exchangers, see SekuliC (1990b).

THERMODYNAMIC IRREVERSIBILITY AND TEMPERATURE CROSS PHENOMENA

765

(1 1.49)

Note that Eq. (1 1.49) is identical to the corresponding relationships in Eq. (3.114). Thus, the number of heat transfer units at the maximum irreversibility in a heat exchanger is exactly the same as the limiting value of NTU (=NTU*) at the onset of an external temperature cross. At that operating point, the outlet temperatures of both fluids are equal. Hence, beyond this NTU > NTU*, there will be a temperature cross, and the hotfluid outlet temperature will become lower than the cold-fluid outlet temperature. As defined in Sections 3.2.3 and 3.6.1.2, the temperature cross derives its name from fictitious or actual crossing of the temperature distributions of the hot and cold fluids in an exchanger. If there is no actual crossing of hot- and cold-fluid temperature distributions within an exchanger and Tc,o> Th,o,we refer to it as an external temperature cross as found in the temperature distributions of a counterflow exchanger in Example 3.2 or the 1-2 TEMA E shell-and-tube heat exchanger of Fig. 3 . 1 7 ~ for the high-NTU case (the solid lines). We will call it an internal temperature cross if there is an actual crossing of hot- and cold-fluid temperature distributions within an exchanger. There are two possibilities for the internal temperature cross: (1) Tc,o> Th,o(as in the 1-2 TEMA E exchanger of Fig. 3.17b at high NTU, or (2) Tc,o,~ocal > Th,o~ocal, where the subscript local means one of the multiple outlets on one or both fluid sides of an exchanger (see the temperature distributions in Fig. 11.7 for the 1-2 TEMA J shell-and-tube heat exchanger of Fig. 11.6). Note that there can be an external temperature cross without an internal temperature cross (such as in a counterflow exchanger, as shown in the temperature distribution of Example 3.2). Let us discuss the implications of the external and internal temperature crosses further in the following two subsections.

11.4.2 External Temperature Cross and Fluid Mixing Analogy


Results obtained in Section 11.4.1 [i.e., an interpretation of the physical meaning of Eq. (1 1.49), and the relation between the equalization of outlet temperatures and maximum entropy generation] lead to yet another interesting analogy. This analogy can be explained by studying Eq. (1 1.49) in a dimensional form. First, let us reconfirm (see also Section 3.6.1.2) the statement of equality of the outlet temperatures at the operating point indicated by Eq. (1 1.49). By invoking definitions of exchanger effectiveness [Eq. (3.44)] and heat capacity rate ratio [Eq. (3.56)], Eq. (11.49) can be rewritten as follows:
(1 1.50~)

Simplifying this equation leads to


TI>O = T2, O = To

(1 1S o b )

So Eq. (1 1.50) confirms that Eq. (11.49) corresponds to an equality of outlet temperatures of the two fluids of a heat exchanger. Now, using the conclusion reached by Eq. (llSO), let us rewrite Eq. (11.49) this time keeping the heat capacity

766

THERMODYNAMIC MODELING AND ANALYSIS

rate ratio in its explicit form:


(1 1.51a)

Therefore,

Equation (1 1.51) clearly indicates an identical result as before, but this time it can be interpreted as having been obtained for a quite different physical situation of adiabatic mixing of the two fluid streams (with heat capacity rates C1 and C2 and inlet temperaSuch imaginary mixing will lead to the outlet temperature To for tures TI,,and T2,,). the mixture of the two given streams (C, C2). From thermodynamics, we know that adiabatic mixing leads to total destruction of the available thermal energy potential (implied by the difference of the temperatures of the fluids) that exists at the onset of mixing. Thus, this process must be characterized by maximum entropy generation. In conclusion, there is an analogy between a heat exchange process at the operating point that corresponds to the condition of an equalization of outlet temperatures and the adiabatic mixing process of the same two fluids. This analogy demonstrates that entropy generation in a heat exchanger at that operating point must be the maximum possible. This constitutes an additional physical explanation of the thermodynamic significance of an external temperature-cross operating point in a heat exchanger. A clear thermodynamic meaning of the result given by Eq. (1 1.49) can easily be confirmed by comparing Eqs. (1 1.28) and (1 1.42), that is, comparing the entropy generations obtained for two completely different processes: (1) a heat transfer process in a heat exchanger characterized with equal outlet temperatures of the fluids involved = T2,0= To: see Eq. (1 1.28)], and (2) a mixing process of the two fluids [i.e., [T, = T,,,,J = 1, 2; see Eq. (11.42)]. The entropy generation rates for two physically quite different processes are found to be identical. Analysis provided so far shows clearly how the facts related to the detrimental influence of fluid mixing on heat exchanger performance fit nicely into the consistent thermodynamic picture. Hence, the results for heat exchanger effectiveness presented in Chapter 3 (Tables 3.3 and 3.6) have deeper physical explanations. Let us now show how this kind of thermodynamic analysis can be used to understand the behavior of a relatively complex heat exchanger flow arrangement. Moreover, we demonstrate why such analysis has importance for practical design considerations.

11.4.3 Thermodynamic Analysis for 1-2 TEMA J Shell-and-Tube Heat Exchanger

It has been emphasized (see Section 3.6.1.2) that contrary to a general design requirement to transfer heat only from one fluid to the other, and not vice versa, in some heat exchangers, reverse heat transfer takes place. For example, consider the 1-2 TEMA J shell-and-tube heat exchanger of Fig. 1 1.6. The existence of a temperature cross leads to a situation in which an addition of more surface area in the second tube pass does not contribute a significant increase in heat transfer because of reverse heat transfer taking place in the second pass. Note that we reached this conclusion in Section 3.6.1.2 based on assumed (i.e., at that point not analytically derived) temperature distributions. In this

THERMODYNAMIC IRREVERSIBILITY AND TEMPERATURE CROSS PHENOMENA

767

T 1 , O

FIGURE 11.6 Schematic of a 1-2 TEMA J shell-and-tube heat exchanger.

chapter (see Section 11.4.2), we provided a thermodynamic interpretation of this undesirable phenomenon through temperature distributions. We elaborated in detail how one can determine temperature distributions and subsequently evaluate the influence of local temperature differences and a mixing process on heat exchanger performance. Let us now show how the internal temperature cross can cause peculiar behavior in the P1-NTUIor E-NTU results. For a given heat capacity rate ratio, with increasing NTUI, the temperature effectiveness PI reaches a maximum beyond which an increase in NTUl causes P I to decrease rather than increase monotonically as may be expected.+This behavior is illustrated in Fig. 3.16. Figures 11.7 and 1 1.8 summarize the results of a thermodynamic analysis of this exchanger. This analysis includes both the first and second laws of thermodynamics. The approach is straightforward. It starts with a determination of temperature and heat transfer rate distributions obtained through an application of the first law of thermodynamics (Figs. 11.7 and 1 1 . 8 ~ ; model formulations in Problem 11.1). Subsequently, entropy generation is determined utilizing both the first and second laws of thermodynamics [see Section 11.3.1 and Eq. (11.33)]. Figure 11.7 shows three temperature distributions for NTU, = 0.87, 1.83, and 5.0 for R 1 = 2. Figure 11.8 provides data regarding the corresponding distribution of dimensionless heat transfer rates [the total, and those determined at different tube sections (a, b, c, and d) and shell zones (A and B) of the Fig. 11.6 exchanger] and entropy generations. To demonstrate this analysis, let us consider a numerical example from the results of Kmecko (1998).

Example 11.4 For a 1-2 TEMA J shell-and-tube heat exchanger, establish the number of temperature crosses and explain the meaning of the existence of a maximum effective'There are also different and unexpected PI vs. NTUl curve behaviors for n-pass n-pass ( n > 2) plate heat exchangers and complex flow arrangements (Shah and Skiepko, 2002).

768

THERMODYNAMIC MODELING AND ANALYSIS

0.6.

-0.4

-0.2

0.2

0.4

5
(C)

FIGURE 11.7 Temperature distributions in a 1-2 TEMA J shell-and-tube heat exchanger; heat capacity rate ratio R I = 2: (a) NTUl = 0.87; (b) NTUl = 1.83; (c) NTUl = 5.00. Note that the scale at the abscissa is extended slightly beyond f0.5 (physical terminal ends) to show clearly the ends of the curves separate from the ordinates. (From Kmecko, 1998.)

<

ness at finite NTUs. The range of operating points under consideration is defined with R I = 2 and a range of NTU values between NTUl = 0.87 and 5.0. The inlet temperature ratio is equal to 2. SOLUTION
Problem Data and Schematic: The schematic of a 1-2 TEMA J shell-and-tube heat exchanger is given in Fig. 11.6. The heat capacity rate ratio R 1= 2. The range of NTU values is 0.87 to 5.0. The inlet temperature ratio 1 9 = 2.0. Determine: The number of temperature crosses for this exchanger and explain the existence of maximum effectiveness at finite NTUs.

THERMODYNAMIC IRREVERSIBILITY AND TEMPERATURE CROSS PHENOMENA

769

0.4

0.2
%

-0.2

101214

NTU,

FIGURE 11.8 (a) Heat transfer rate (R, = 2), and (6) dimensionless entropy generation rate (R, = 1 and 2, 29 = 2) in a 1-2 TEMA J shell-and-tube heat exchanger. Note that the NTUl axis is extended slightly beyond 0 to show clearly the ends of the curves separate from they axis. (From Kmecko, 1998.)

Assumptions: The assumptions are those listed in Section 3.2.1.

Analysis: The solution of the mathematical model of this heat exchanger (for the model formulation, study Problem 11.l) provides temperature distributions for both shell and tube fluids. The solution method may be the Laplace transforms technique, as discussed earlier in this chapter, or any other pertinent method as mentioned in Section 3.1 1. These distributions are given in graphical form in Fig. 11.7 for a fixed heat capacity rate ratio (R,= 2) and three values of NTUI. The NTUl values correspond to the operating points for the following three specific cases: (1) the equalization of exit temperatures at small NTU (the first external temperature cross at NTUl = 0.87 at the exchanger outlets of two fluids (TI,, = T2,,), (2) the maximum temperature effectiveness at NTUl = 1.83), and (3) the equalization of exit temperatures at a large NTU value (the second external temperature cross at NTUl = 5.0 at the exchanger outlet, TI,, = T2,,).In all three cases, we have internal temperature crosses. corresponds to an operating point The temperature distribution in Fig. 1 1 . 7 ~ NTUl = 0.87 and R1= 2.0. The fluid 2 (tube fluid) dimensionless temperature decreases along both tube passes (tube segments d-b-a-c in Fig. 11.6 and 02,d -+ O2,b + e2,, + e2,c in Fig. 1 1 . 7 ~ reaching ) the fluid exit with the dimensionless temperature e2,, (its corresponding dimensional value is Tz,,). The outlet temperature of fluid 2 at that operating point is exactly equal to the mixed mean temperature of fluid 1 obtained after mixing both exit streams of fluid 1 [i.e., T , , , = (T;,, T;:,)/2,or in dimensionless = (el,*,, C31,B,o)/2].To emphasize, 0 2 , , = el,, or T2,, = T I , , for form el,, NTUl = 0.87 and R, = 2 for Fig. 1 1 . 7 ~Note . that fluid 1 (shell fluid) dimensionless temperature changes (increases) along both shell zones (A and B) following the distributions presented in Fig. 1 1 . 7 (@,,A ~ and e I , ~ Therefore, ). this operating point corresponds to an equalization of exit temperatures, that is, an onset of an external temperature cross (ETC). In addition, an internal (or actual) temperature cross (ITC) is present in zone B at

Next Page

770

THERMODYNAMIC MODELING AND ANALYSIS

the ITC point = Note that the temperature effectiveness of fluid 1 is P I = el,, and P2 can then bc calculated from P2 = P IR I. If N T U l is increased from 0.87 to 1.83, a new operating point will be reached. Under that condition, temperature distributions will become as given in Fig. 11.7b. This figure in the second reveals a peculiar increase of the fluid 2 dimensionless temperature (02,,) part of the tube segment c: instead of decreasing, the original hot tube fluid 2 dimensionless (or dimensional) temperature increases. This is due to reverse heat transfer taking place to the left of the ITC point in Fig. 11.7b. As expected, the resulting exit temperature of fluid 1 is higher than that for the operating point presented in Fig. 1 1 . 7 ~This . means that the temperature effectiveness of fluid 1 has increased from (= Ol,o,a)to (the subscripts a and b denote the cases from Fig. 11.7a and b, with (= NTU = 0.87 and 1.83, respectively). If we increase the number of transfer units even further, say from 1.83 to 5.0 (see Fig. 11.7c), the increase in the fluid 2 dimensionless temperature 02,c in the tube segment c becomes very pronounced (see, e.g., 02,c at [ = -0.5 in Fig. 11.7~). The shell fluid temperature still continues to increase in both shell zones, but much more so in zone B, O I B in Fig. 1 1 . 7 ~ . However, of Fig. 11.7b. As a result, the [= (Ol,A,o O1,,,,)/2] of Fig. 1 1 . 7 ~ is less than temperature effectiveness of fluid 1 will become smaller than the corresponding value of Fig. 1 1.7b. This means that with an increase of N T U l starting from 1.83, the temperature effectiveness decreases from its maximum value at N T U l = 1.83 and R I = 2.0, as shown in Fig. 11.8a, in the q* (= P I )vs. NTUl curve. T o understand this peculiar behavior, let us now consider the distribution of heat transfer rates within the heat exchanger in various shell zones and tube segments, as presented in Fig. 1 1 . 8 ~ . that with an increase in N T U I , an initial It is interesting to notice from Fig. 1 1 . 8 ~ - TI.[)] = P I , is increase in total dimensionless heat transfer rate, q* = q/[(hcP)l(T2,i followed by a maximum at NTU, = 1.83 and decrease at higher N T U I . Adding more heat transfer area beyond that for the maximum effectiveness decreases the heat transfer rate exchanged between the two fluids. How could that be possible? The components of this heat transfer rate are presented in the same diagram (Fig. 11.80) for both the shell fluid (zones A and B) and the tube fluid (tube segments a to d) having appropriate subscripts with q*. From this figure, it is clear that only the tube segment d (the inlet segment of the first tube pass) contributes to the heat exchanger performance by its positive dimensionless heat transfer rate slope in the direction of increasing N T U I . The heat transfer contribution in tube segments a and b (and correspondingly, in the shell zone A) is diminishing rapidly, while the heat transfer in tube segment c becomes reversed and is increasing in the negative direction. Consequently, a large portion of the heat exchanger actually fails to contribute to the original design goal of increasing q with P I at large NTUl values, beyond NTU, = 1.83 for this case!

Di.~cussionand Comments: The behavior of the analyzed heat exchanger can be interpreted in terms of entropy generation. For this exchanger, $,,can be calculated using Eq. (1 1.33) for 8 = Tl,,/T2,, = 2.0 and the P-NTU formula of Eq. (111.1 1) of Table 3.6 for given NTUl and R I = 1 and 2 as shown in Fig. 11.8b. For this exchanger, the entropy generation features two maximums and one local minimum. The two equalizations of outlet temperatures (ETCs, one at smaller and the other at larger N T U J correspond to two maximums in entropy generation for the case of R I = 1 and 2. In this situation, the first is at NTUl = 0.87 and the second at NTUl = 5.0, as shown in Fig. 11.8b. At the operating point characterized by the maximum temperature effectiveness, the heat

Potrebbero piacerti anche