You are on page 1of 231

Process Design:

The Perpetuation of Wrong Designs

The author hereby consents to that this book may be downloaded, copied and distributed
without limit and gratis.

Juon C Wah
Calgary
Fall 2014
Table of Contents

Chapter 1: Overhead Pressure Control for Fractionation Towers 1

Chapter 2: Steam as a Heating Medium 20

Chapter 3: Separators 32

Chapter 4: Oil Skimming in a Horizontal Separator 64

Chapter 5: Shell and Tube Heat Exchangers 78

Chapter 6: Thermosiphon Reboilers 108

Chapter 7: Thermodynamics – An Overview 123

Chapter 8: Refrigeration 162

Appendix I: Calculation Routine for Steam Desuperheater Control 188

Appendix II: Principle of Launching with Angular Momentum 190

Appendix III: Design and Sizing Considerations for a Liquid Launcher 192

Appendix IV: Bernoulli Equation from the First Law 201

Appendix V: Design of Three Phase Separator with Oil Skimming 205

Appendix VI: Pinch Temperature and Exchanger Duty Shortfall 216

Appendix VII: LMTD Correction 220

Appendix VIII: Thermodynamic Relations 222

Appendix IX: Batch Boiler 225

References 227
PREFACE

In more than three decades spent on process design and process operation in the
hydrocarbon processing industry, the author has noticed quite a number of wrong
designs. By wrong designs, it is meant that equipment or process units built according to
them will not work. The causes of these wrong designs are first and foremost the
misconception or misunderstanding of the physical principles governing the workings of
the process. The failure to work of these designs has been observed in the operation of
actual equipment and process plants built according to them. These wrong designs
perpetuate themselves from one project to another for decades; they are widespread
among the engineering houses and technology vendors everywhere.

Analyses of these wrong designs are presented. The factors, both technical and
organizational, that may have contributed to their perpetuation are discussed. Of course,
for each of these wrong designs a right one that does work is presented.

The chapters in this book are autonomous; they may be read in any order.

He and She

In this book, when a third person singular is referred to, the pronoun he (or its derivatives
his and him, as the case may be) is used. This use is made for the sake of economy and to
avoid the rather clumsy constructions such as: “…he or she…”, “…his or her…”, or
“…him or her…”; or “…he/she…”, “…him/her…” No offence to gender equality is
intended in this exclusive use of the third person singular in the masculine form. The
author is delighted to note that over the years there has been a marked increase in the
number of women engaged in the engineering profession.
ACKNOWLEDGEMENT

I am deeply grateful to my life-long friend Peter C Flynn, Professor Emeritus, University


of Alberta, for his tireless effort and long hours he spent on editing the manuscript. His
contributions and advice have been invaluable.

Any errors and omissions that may occur in this book in its present form remain entirely
my responsibility.
Chapter 1:

OVERHEAD PRESSURE CONTROL FOR FRACTIONATION TOWERS

1.1 Introduction

For a fractionation tower to function properly it is of paramount importance that its operating
overhead pressure be held constant at a desired value. This desired value of the operating
pressure is usually determined by design to meet the fractionation objectives of the tower. Once
so determined, this pressure as an operating parameter usually remains unchanged throughout the
operation of the tower.

The methods of control of this operating pressure, i.e. keeping it constant, depend largely on
whether or not a continuous stream of vapour product is withdrawn from the tower overhead. In
the case of the presence of a continuous vapour product stream in the overhead, the tower either
has either a partial condenser or no condenser at all. In a partial condenser, only part of the total
tower overhead vapour leaving the top tray of the tower is condensed; that part which is not
condensed is withdrawn as the overhead vapour product. Part of the liquid produced in the
condenser is returned to the tower as reflux, the remainder withdrawn as overhead liquid product.
In some applications, the liquid produced by the partial condenser is returned in its entirety as
reflux to the tower – a set-up known as total reflux - in which case the tower has no overhead
liquid product. In the case of a tower with no overhead condenser (an absorber, for example), all
vapour leaving the top tray of the tower is withdrawn as vapour product.

In the case of zero overhead vapour product, all the vapour from the top tray of the tower is
condensed. The tower is then said to operate at total condensation. Part of the liquid produced in
the condenser is returned to the tower as reflux and the remaining withdrawn as overhead liquid
product.

When a steady stream of overhead vapour product is withdrawn, the control schemes for tower
overhead pressure are usually well understood and devised correctly. However, in the case of
total condensation, i.e. no vapour is withdrawn as an overhead product, the phenomenon of the
overhead pressure is often misunderstood; this misunderstanding leads to ill-conceived control
schemes that do not work, among which figures prominently the “hot vapour bypass” control.

In Section 1.2 we present a brief review of the pressure control schemes for the fractionation
tower that makes a steady stream of overhead vapour product. In Section 1.3 we discuss at some
length the hot vapour bypass method, a method that does not work for the control the overhead
pressure of a fractionation tower with total condensation, and we also analyze the common
causes of the misconception leading to the use of this dysfunctional scheme. In Section 1.4 we
present the semi-flooded tube method, a method that works for the pressure control of a
fractionation tower with total condensation.

1
1.2 Overhead Pressure Control with a Steady Overhead Vapour Product Stream

Figure 1.1 (a) to (d) show the pressure control schemes used successfully for a variety of
fractionation applications. In all these schemes, the underlying principle is the same: the desired
pressure in the tower overhead is obtained by what is commonly known as “back pressure”
control.

PC

PC
Vapour
product Vapour
product
Reflux
Feed

Liquid product ≥ 0

(a) Stripper or absorber with no (b) Refluxed column


rectifying section and no reflux
SP Position
SP Speed PC
PC
controller
controller Vapour
product
Vapour
product Reflux
Reflux Suction
Loader

Liquid product ≥ 0 Liquid product ≥ 0

(c ) Refluxed column - vapour product to (d) Refluxed column - vapour product


a compressor with variable speed to a fixed speed compressor with
control suction loader

Figure 1.1:
Pressure control for fractionation tower having an overhead vapour product stream

This back pressure control scheme is easy to understand. It entails varying the restriction against
which the vapour product is made to flow out of the tower overhead. For a given mass flow rate
of the vapour, the higher the restriction the higher is this “back pressure”, i.e. the pressure
established upstream of the restriction required to drive the same vapour mass flow rate out
against this restriction. Figure 1.1 (a) shows a scheme for a tower with no condenser (hence no
reflux, either); all the vapour leaving the top tray of the tower is the overhead vapour product.
The variable restriction to flow here is provided by the control valve; similarly in Figure 1.1 (b),
which represents a refluxed tower with a partial condenser. Figure 1.1 (c) and (d) represent
applications where the tower pressure is not high enough to deliver the vapour product to its
destination; a compressor is used to enable such delivery. With a compressor in the scheme, the

2
varying of restriction to flow can be achieved in two ways. In Figure 1.1 (c) the restriction is
varied by varying the speed of the compressor: for a given vapour product mass flow rate, the
compressor speeds up (i.e. decreases the flow restriction to its suction) to draw down the tower
overhead pressure; it slows down (i.e. increases flow restriction) to back up a higher pressure in
the tower overhead. In Figure 1.1 (d), a fixed speed compressor is used. Here, the position of a
mechanism in the so-called “suction loader” is manipulated to vary the restriction to flow. The
suction loader is effectively a valve, a device that varies the flow area to the suction of the
compressor; and it usually comes as an integral part of the compressor suction. For a compressor
without an integral suction loader, an extraneous damper or adjustable vane may be used to
provide the means of varying the restriction to flow to the compressor suction.

1.2.1 Partial Condensation: Constant Condenser Duty

It is clear in the control schemes shown in Figure 1.1 that the desired pressure in the tower
overhead is obtained by varying the flow restriction against which the overhead vapour product
is made to exit the overhead. However, a common misconception is that the overhead pressure is
controlled by varying the rate of heat removed in the condenser thereby varying the rate at which
vapour is condensed. The reasoning, according to this misconception, goes like this: when the
pressure is too high, more heat is removed by the condenser to condense more vapour in order
bring the pressure down to the desired value. Conversely, when the overhead pressure is too low,
the heat removal by the condenser is reduced to reduce the rate of condensation (vapour
disappearance) to bring the pressure back up to the desired value. Many ill-conceived and
dysfunctional control schemes are based implicitly or explicitly on this common misconception.
See Buzzeta et al.(4).

In all fractionation applications (partial or total overhead condensation), the condenser duty is a
constant and meant to be kept constant during operation. The condenser is sized, for a given
reflux rate, for this full unvaried duty that caters for the highest demand. This highest demand is
constituted by a combination of throughput and other requirements such as product quality. The
question naturally arises as to what the operator should do if the tower operates below design; for
example at lower than design throughput. Should the condenser duty be reduced accordingly? If
not, would the heat removal at full design rate by the condenser be excessive and cause the
overhead products, both liquid and vapour, to go off specification on cut points and rates, or
cause some other problems? The answer is no, the condenser duty should not be reduced
accordingly; holding the condenser duty constant does not result in excessive heat removal and
wrong product cut points or product rates or quality, or some other undesirable consequences.
This is so because the reboiler and other sources of vapour generation for the tower are held
constant as well. In fact it is desirable, and sometimes even necessary, to maintain constant the
reboiler and condenser duties, as well as other sources of vapour and liquid generations internal
to or associated with the tower, in the face of changing fractionation throughput or product

3
requirements. By maintaining these duties constant, the tray loadings in the tower are thereby
kept nearly constant. A tray loading refers the rates of vapour flow and liquid flow through a
fractionating tray. The loading of a tray determines the adequacy of that tray to provide the
liquid-vapour contact necessary for mass transfer (fractionation). While a tray may be designed
with a tolerance to some variation in loadings (so that the tray does weep when run at lower
vapour rates, for example), the effectiveness of mass transfer is usually optimal at or near the
design tray loadings, loadings in terms of liquid and vapour flows rates through the tray as well
as the ratio of these flow rates. In other words, for the operation of a fractionation tower, the
internal vapour and liquid traffics (i.e. tray loadings) are usually kept nearly constant, regardless
of fractionation throughput or product requirements (within reasonable variation, of course). This
near constancy of tray loadings is achieved, among other measures, by keeping constant the heat
removal in the condenser as well as heat input in the reboiler. The “other measures” mentioned
in the preceding statement refer to pump-around heat removal rate and side stream stripping rate
for complex fractionation applications such as a crude tower in a petroleum refinery.

The reflux rate of a fractionation tower has to be first and foremost adequate for the fractionation
(separation) required. This reflux rate, adequate to meet the fractionation objective, must also be
adequate to provide robust vapour and liquid loadings of the trays – robust in this context means
that any change in tray loadings in response to feed fluctuation is such that the trays still perform
near the optimum. It is seen that the greater the reflux rate, the greater is the robustness of the
tray loadings. The range of loadings in which a tray remains robust depends on the type of trays.
The sieve tray has a narrower range than the valve tray, whereas the bubble cap tray has a much
wider range of any of the two. This means that the bubble cap tray can remain robust with a
lower reflux rate than the sieve or valve tray. The cost of trays varies greatly according to their
type: the cost of the sieve tray is lowest and the bubble cap tray the highest, with the valve tray
somewhere in between. In selecting the reflux rate, which in turn determines the condenser and
reboiler duties, both the fractionation requirement and the robustness of tray loadings must be
considered, in the light of the type of trays to be used.

It should be noted that in most if not all fractionation applications, the overhead condenser is not
equipped with a control mechanism, whether manual or automatic, that allows the condenser
duty to be adjusted as an operating variable. This is so because, for reasons discussed above, the
condenser is meant to run at constant duty. (Occasionally, one may see ill-conceived control
schemes for a fractionation overhead that include features to vary the duty of the condenser, such
features as an adjustable coolant flow for a shell and tube condenser.) However, it should noted
that an air-cooled condenser used in a region with drastic change in the ambient temperature is
provided with control schemes – such as adjustable fan speed or louver position – to maintain a
nearly constant condenser duty. Any small amount of over-cooling resulting in a higher reflux
rate can be compensated by increased reboiler duty to maintain the desired product rates and
qualities.

4
Furthermore, within the pressure variation about the target (set point) expected of the tower
overhead pressure control, the heat of condensation (hence condenser duty) is practically
constant (see Figure 1.2). This further underscores the fact that overhead pressure of an operating
fractionation tower is not controlled by varying the heat removal rate (duty) of the condenser.
The difference in condenser duty between P 1 and P 2 in Figure 1.2 is too small to provide enough
resolution for control.

B1 = ∆H C , heat of condensation at P 1
P set
P1 P2 = condenser duty
B 2 = ∆H C , heat of condensation at P 2
B2 = condenser duty
B1 M
M1 B 1≈ B 2
Enthalpy, H

A2
A1 A 1/ B 1, A 2/ B 2 = vapour / liquid ratios
for partial condensation at P 1, P 2;
these ratios are practically constant

Pressure, P
Figure 1.2:
Phase envelop (P -H ); partial condensation at P 1 and P 2 ; within
the variation between P 1 and P 2 about the target P set , the
condenser duty (heat of condensation, measured by B 1 and B 2)
is practically constant

1.3 Overhead Pressure Control with Total Condensation: The Hot Vapour Bypass Scheme

Much misconception has been noted when it comes to controlling the overhead pressure of a
tower that operates with total condensation and consequently makes no vapour product. Unlike
the case of partial condensation, here we simply do not have an exiting vapour product stream to
manipulate for the purpose of pressure control. For this case the hot vapour bypass method (or
technique) for overhead pressure control is the most prevalent scheme offered by the engineering
houses and technology vendors. Quite often this is the only method offered. This hot vapour
bypass method, in various guises, has made its appearance in many installations as well as in the
literature. Many scholarly or technical papers have been written about fractionation tower control
on the premise that this scheme works: Friedman (7) and Buzzeta et al (4), for example.
Nonetheless, the dysfunction of the hot vapour bypass method has been recognized and
discussed to considerable extent by various authors: Sloley (2), Chin (5) and Kister (6). The most
succinct statement is one made by Sloley: It does not work. We discuss below this method and
the misconceptions that engender it.

5
1.3.1 The Commercial Hot Vapour Bypass Scheme

Figure 1.3 shows a variety of this hot vapour bypass scheme. The main difference in the variety
is where the pressure is sensed: upstream of the condenser or downstream of it, as shown in
Figure 1.3 (a) and (b). In (c), a split range of the pressure controller output is used so that, as
intended by the designer, if the hot vapour bypass cannot keep the pressure down to the desired
value, the excess vapour can be vented off to flare in order to draw down the tower pressure, as a
last resort! In short, the scheme in (c) is designed to fail. This is so because when the overhead
vapour is vented to flare the overhead liquid product (and most probably the other fractionation
products as well) will go off-specification. As far as fractionation is concerned, this constitutes a
failed operation, and it becomes immaterial whether or not the tower pressure is still under
control in the event of such a failed operation. A failed operation must be terminated.

Over the years the present author has talked with proponents and believers of the hot vapour
bypass method and studied their published discourses to find out how and why they think it
works. Summarized below in general terms (Section 1.3.4) are their arguments and reasoning.
But let’s first look at the anatomy of the hot vapour bypass scheme and the constancy of
condenser duty for total condensation.

PC
Hot vapour
Hot vapour
bypass
bypass

Reflux
Reflux
PC

Liquid product ≥ 0
Liquid product ≥ 0
(b) Refluxed column -
(a) Refluxed column -
hot vapour bypass control - a variation
hot vapour bypass control
Split range
PC
Hot vapour
bypass
Vapour to flare
NNF (normally no flow )
Reflux

Liquid product ≥ 0
(c ) Refluxed column -
hot vapour bypass control - a variation
Figure 1.3:
Pressure control for tower with total condensation (no vapour product);
the dysfunctional "hot vapour bypass technique"
6
1.3.2 Anatomy of the Hot Vapour Bypass Scheme: Why It Does Not Work

Figure 1.4 shows the anatomy of the hot vapour bypass scheme. Let’s first look at the system
defined by the H&MB (heat and material balance) envelope in the figure. We designate
everything inside the envelope as the system. The only material stream entering the system is the
saturated vapour from the top tray of the fractionation tower; the only material stream leaving is
liquid from the reflux drum. This means total condensation has to occur in the system; all vapour
entering must first become liquid and exits the system as liquid. All this becoming occurs on the
tube surface of the condenser. The heat of condensation QC is the heat leaving the system, i.e. it is
the heat transferred from the condensing vapour and carried away by the coolant flowing through
the condenser. It is important to note that this is the only heat transaction between the system and
the surroundings – barring negligible heat loss from the piping and equipment surface.

Vapour (from tow er top) QC = condenser duty (heat


Surroundings removed by the coolant)
PT
QC System
PC H&MB (heat & material balance)
Hot vapour
bypass envelope
Coolant OUT
Vapour spaces in condenser and
reflux drum communicate; so that
PC PT = PD ≅ PC

Coolant IN
PT and PD are equalized due to
Condenser the open valve on the hot vapour
bypass line. Flow in the line stalls;
the hot vapour bypass does not
PD
bypass!
Reflux drum
Vapour space in reflux drum is a
dead end
Liquid
Figure 1.4 : (product & reflux)

Anatomy of the hot vapour bypass scheme; the hot vapour bypass does not
actually bypass

The vapour space in the reflux drum is a dead end for the vapour entering the system; as all
vapour entering the system must get condensed in the condenser; there is no exit route for the
vapour to leave the system, once it has entered it. (The preceding statement also applies to Figure
1.3 (c), since the control valve to vent vapour from the reflux to flare is intended by design to be
normally closed. (Venting of the overhead vapour to flare constitutes a failed operation and lies
outside the purview of the analysis here, which analysis is on how the control scheme supposed
work during normal operation.) During normal operation the control valve on the hot vapour
bypass line must assume a certain partially open position in order to remain modulating; a fully
open or fully closed control valve does not modulate and hence does not control. When the
control valve is open, to whatever extent it may be open, all that happens is that the pressure P D

7
in the vapour space of the reflux drum equalizes with the condenser inlet pressure P T , which is
effectively the tower pressure. Once these two pressures are equalized, vapour in the hot vapour
bypass line stagnates; the hot vapour bypass simply has no place to bypass into, the vapour space
in the reflux drum being a dead end. Varying the opening of the control valve will have no effect
on the tower pressure PT or on the stagnation in the hot vapour bypass line. In other words, the
hot vapour bypass, on which this control scheme is based, does not bypass and therefore this
scheme cannot possibly work as intended by its designer.

1.3.3 Total Condensation: Constant Condenser Duty

As in the case of fractionation applications making an overhead vapour product (see Section
1.2.1), here too a prevalent misconception is that the pressure control is achieved by varying the
rate of condensation. This is in spite of the obvious fact that we have here total hence perforce
constant condensation; all vapour entering the system must be condensed. This is the
misconception underlying the dysfunctional hot vapour bypass scheme.

P set
P1 P2

B2
Enthalpy, H

B1
B1 = ∆H C , heat of condensation at P1
B2 = ∆H C , heat of condensation at P2

Pressure, P
Figure 1.5:
Phase envelop (P -H ); total condensation at P1 and P2 ; within the expected
overhead pressure variation between P1 and P2 about the target P set , the
condenser duty, measured by B 1 and B 2 , is practically constant

As in the case of fractionation with partial condensation overhead vapour product (Figure 1.2),
here too, within the expected variation of the operating pressure, the heat of condensation varies
little, as shown in Figure 1.5. This should, again, help dispel the notion that the pressure is
controlled by condensing more or less, or by removing more or less heat. However, an air-cooled
total condenser used in a region of drastic change in ambient temperature is provided with a
control scheme (adjustable fan speed and louver position) to ensure constant condenser duty.
This control is used in conjunction with varying the condenser tube area. On a cold day, the tube
area may have to be flooded by 70 % to maintain the fractionation tower overhead pressure.
When this occurs the fan speed should be stepped down to reduce the air flow so that the flooded

8
tube area assumes a good control level (20 to 30 % flooded). (See Section 1.4 on the pressure
control method for a total condenser.)

1.3.4 Why Do People Think the Hot Vapour Bypass Scheme Works?

There is considerable misconception of the fundamental forces at work here. Summarized below,
into four categories, is the fallacious reasoning articulated by the proponents of the hot vapour
bypass control scheme:

Fallacy 1: Pressure Control by Bleeding Off Excess Vapour

This fallacy might have been inspired by the observation that the pressure prevailing in an
enclosed volume can be adjusted by bleeding off the excess gas or vapour held in it, much in the
same way as we would reduce the pressure of an over-inflated tire to a desired value by bleeding
off the excess air in it. The reasoning goes further here: since vapour from the tower enters the
system continuously, the tower pressure P T can be controlled to any desired value by bleeding
off continuously the correct amount of vapour through the hot vapour bypass line. In this
reasoning, it is not recognized that the vapour space in the reflux drum is not an extraneous sink
(in the manner that the atmosphere is an extraneous sink into which we bleed the excess air
captured in an over-inflated tire), but a dead end that forms a part of the vapour space of the
system itself; hence there can be no bleeding off of the excess vapour. (See Figure 1.4.)

Fallacy 2: Removing Excess Vapour by Condensing It in the Reflux Drum

Here again the reflux drum is not recognized as a dead end but an infinite sink for the hot vapour
bypass that flows into it. It is believed that the hot vapour, upon meeting the cold liquid in the
drum, condenses in it. It is argued that this removal of vapour by condensing lowers the pressure.
As the vapour flow (from the tower) into system is continuous, the flow of vapour bypass into
the reflux drum and its subsequent condensation in the liquid there is also continuous. The rate of
the hot vapour bypass is adjusted by the control valve in the hot vapour bypass line so that right
amount of hot vapour is let through and allowed to disappear from the vapour space by
condensing into the cold liquid in the reflux drum, continuously, to maintain the tower pressure
P T at any desired value.

Implicit in this reasoning is the assumption that the liquid in the drum is always sufficiently sub-
cooled so that the heat of condensation imparted to it by the condensing hot vapour bypass is all
absorbed as sensible heat by the liquid in the drum; this absorption of the heat, it is assumed,
causes the liquid temperature to rise but not enough to cause it to boil and generate vapour. Also

9
implicit is the assumption that in the reflux drum vapour-liquid contact area is always sufficient
to cater to the required rate of vapour condensing into the liquid. Both these assumptions are
untenable. The liquid (condensate) that enters reflux drum from the condenser is to all intents
and purposes saturated; any subcooling that this liquid may receive during its residence in the
condenser is incidental and minimal. The condenser is well drained, which means all tube
surface is exposed to the condensing vapour and no tubes are submerged in the condensate. The
condensate dropping off the tube surface is in full contact with the condensing vapour at all times
during its residence in the condenser; the liquid collected in the reflux drum is therefore perforce
saturated, i.e. liquid at its bubble point. In addition the reflux drum is a rather poor device to
promote vapour-liquid contact; only the liquid surface in the drum and the surface of the streaks
of liquid falling into the drum are accessible to the vapour.

The hot vapour bypass is saturated since it is the same vapour that leaves the top tray of the
tower. This vapour is therefore at its dew point. As this saturated vapour comes into contact with
the colder but saturated liquid in the reflux drum, it condenses by yielding the latent heat of
condensation to the liquid. Because the liquid is at its bubble point, the heat of condensation
imparted to it in turn causes it boil, generating vapour to replenish the vapour space of the reflux
drum. This is clearly a zero-sum game, with no net reduction of the vapour population resident in
the vapour space, as the removal of vapour by the hot vapour bypass condensing is countervailed
by the generation of vapour by the liquid boiling. (To believe that the reflux drum can act as a
sink into which the “excess vapour” can be made to disappear is somewhat akin to the belief that
a person can lift himself off the floor by pulling hard enough on his bootstraps.)

This zero-sum game can also be seen by looking at the H&MB envelop shown in Figure 1.4. The
only heat stream leaving the system is the heat of condensation absorbed and carried away by the
coolant. The heat Q C so removed corresponds to the rate of vapour condensed, which in the case
of total condensation is equal to the rate of vapour entering the system (from the tower). Any
interchange of heat and material within the H&MB envelope, such as between the hot vapour
and the cold drum liquid, cannot affect Q C or, consequently, the amount vapour condensed,
which we already know has to be total and constant, no more no less.

To demonstrate this zero-sum game, let’s give the proponents of this fallacy the benefit of the
doubt and grant that the liquid enters the reflux drum subcooled by 1 °C. For this purpose, we
use: 110 kJ/kg-mol/K for the specific heat of the liquid and 16000 kJ/kg-mol for the heat of
vaporization. (These two quantities correspond approximately to those of light hydrocarbons,
such as those encountered in the overhead of a depropanizer tower, for example.) When a kg-
mole of the hot vapour condenses it imparts 16000 kJ of heat to the liquid; of this less than 1 %
(110 kJ) is used to heat the liquid up to saturation, i.e. to raise the liquid temperature by 1 °C to
reach its bubble point; and the rest of heat (greater than 99 % of 16000 kJ/kg) imparted to liquid
results in generating vapour from the liquid, i.e. boiling it.

10
A hot vapour bypass proponent may be inclined to argue that if the liquid enters the drum a lot
more sub-cooled than by 1 °C, there will be a substantial net vapour removal from the vapour
space to make this hot vapour bypass work. It is argued that more subcooling can be achieved in
the condenser by increasing its duty by, for example, increasing the cooling rate or lowering the
coolant temperature. But as we mentioned before, during its residence in the condenser, the
condensate dropping off the tube surface is in full contact with the condensing vapour at all
times; the liquid collected in the reflux drum is therefore saturated, i.e. liquid at its bubble point.

So, far from being a sink that can remove vapour continuously to regulate the overhead pressure,
the vapour space in the reflux drum turns out to be a dead end.

Fallacy 3: Generating Vapour by Boiling Off the Liquid in the Reflux Drum

Similar to the Fallacy 2, it is believed here that the pressure can be regulated by controlling the
amount (moles) of vapour in the vapour space of the system. However, here the reflux drum
liquid is considered a “vapour source” and not the “vapour sink” on which Fallacy 2 is based.
This is Fallacy 2 in reverse. Here the reasoning goes like this: If the pressure is lower than
desired, it means more vapour needs to be generated to populate the vapour space to bring the
pressure back up. The hot vapour bypass opens to regulate the flow of hot vapour into the reflux
drum. This hot vapour bypass in turn supplies heat to the liquid in the drum to boil it, generating
vapour to replenish the vapour space. As vapour from the tower enters the system continuously,
hot vapour is bypassed continuously into the reflux drum, at a rate regulated to generate the right
amount of vapour to maintain the desired pressure.

Here, as before, the zero-sum game prevails. The hot vapour, being saturated, has first to
condense (loss of vapour) in order to heat up the liquid to boil (gain of vapour). The liquid in the
reflux drum cannot a “vapour source” any more than it can be a “vapour sink”.

As before, we can also see this zero-sum from the prospective of the H&MB (Figure 1.4) as
discussed above.

Fallacy 4: Regulating the Vapour Flow through the Condenser Using the Hot Vapour Bypass

Again, the underlying misconception is that the overhead pressure is controlled by condensing
more or less of the vapour that enters the system from the tower. But the reasoning here is a lot
less sophisticated and less demanding than that in the preceding two fallacies. It does not require
the hot vapour bypass to condense in the reflux drum liquid to deplete vapour; nor does it require
the reflux drum liquid to boil to replenish vapour. It is simply argued here that if the pressure is
lower than desired, “obviously” too much of vapour entering the H&MB envelope is being
condensed; to correct this situation, the hot vapour bypass valve is open to allow more hot

11
vapour to bypass or avoid the condenser; this results in less vapour going through the condenser,
hence less is condensed and the pressure is thereby restored to the set point (desired) value.
Conversely, if the pressure is too high, then more vapour needs to be condensed; the control
valve on hot vapour bypass line moves to restrict the flow, so that more vapour is forced through
condenser to get condensed.

As we know, all vapour entering the system must condense - no more and no less - regardless of
the opening of the control valve on the hot vapour bypass line. As discussed above, all the hot
vapour bypass line does is to equalize the pressure in the vapour space in the reflux drum with
the condenser inlet pressure; this pressure equalization causes the vapour in the hot vapour
bypass line to stagnate.

1.4 Overhead Pressure Control with Total Condensation:

Semi-flooded Tube Method, a Scheme That Works

Figure 1.6 shows a scheme for the pressure control for the tower overhead that has a total
condenser, hence no vapour product. The pressure control is achieved by varying the heat
transfer area exposed to the condensing vapour. This varying of heat transfer area is achieved by
adjusting the level of liquid held in the condenser; as the level is raised, less tube area is exposed
to the condensing vapour, and vice versa. On the exposed tube area condensation occurs; on the
submerged tubes minor subcooling of the liquid takes place. Consequently, of the total condenser
duty Q C greater than 99 % is due to vapour condensing on the exposed tube area.

Here we don’t have a continuous material flow of exiting vapour product whose flow restriction
(valve opening) we can vary in order to back up the pressure. But there is nothing stopping us
from varying the heat transfer area for heat flow! This is precisely what the semi-flooded tube
method is set to accomplish. When pressure falls below the desired value (set point) the liquid
level in the condenser is raised to reduce the heat transfer area exposed to the condensing vapour.
This reduction in area increases the restriction to constant heat flow Q C . The condensation
process responds to this increased restriction by condensing at a higher temperature in order to
drive the same Q C across the now reduced tube area wall; i.e. the temperature of condensation
(or bubble point temperature) is “backed up” due to the reduced heat transfer area. (This is
analogous to backing up pressure by reducing the valve opening to the exiting vapour stream, in
the case of partial condensation of the overhead.) To a higher bubble point temperature
corresponds a higher pressure, namely the pressure at which condensation takes place. The tower
overhead rides on (or “sees”) this bubble pressure P C prevailing in the condenser; in other words
the bubble point pressure determines the tower overhead pressure P T .

12
Pressure equalizing line (no valve )
PT
Inert purge
Coolant (intermittent) PC

PC
Reflux
Coolant

Condensate level
(manipulated variable) PD

PT = PD ≈ PC
Liquid product ≥ 0

Figure 1.6:
Semi-flooded tube method, a scheme that works: total condensation;
zero vapour product; pressure control achieved by varying tube area
exposed to condensing vapour

Conversely, when the overhead pressure is higher than desired, the controller moves to lower the
liquid level in the condenser, thereby increasing the heat transfer area making it less restrictive to
the constant heat flow Q C . The process responds by condensing at a lower temperature, hence a
lower bubble point pressure P C , which pressure, as we have said, is pivotal to the tower
pressure.

The dynamics of the interplay between the heat transfer area and temperature back-up in the
condensation process can be seen from the heat transfer equation: QC = UA∆T ; where Q C is the
constant heat duty; U the overall heat transfer coefficient; A the heat transfer area and ∆T
temperature difference (log-mean) between the condensing vapour and the coolant. When the
tube area A is reduced ∆T must increase to compensate in order to drive the same Q C across the
tube wall; this results in an increase in the condensing temperature (or bubble point temperature),
since the coolant side temperature profile remains the same (at constant duty Q C and constant
coolant flow rate). The overall heat transfer coefficient U remains practically constant within the
range of temperatures and pressures in which condensation is expected to take place.

As noted above the latent heat of condensation is 2 to 3 orders of magnitude greater than the
sensible of subcooling the liquid by 1 °C. Furthermore, the condensation process has a heat
transfer (film) coefficient 2 to 3 orders of magnitude greater than that expected of liquid
subcooling in the condenser. Any subcooling of the liquid obtained with the submerged tube area
is incidental and insignificant. Therefore practically all heat transfer (greater than 99 % of the
condenser duty Q C ) occurs across the tube area exposed to the condensing vapour. It should be
noted that subcooling, to whatever extent it may occur, has no bearing on the overhead pressure
in the pressure control scheme shown here; since subcooling subsequent to condensing does not

13
affect the condensing temperature (bubble point temperature), it cannot affect the bubble point
pressure P C .

1.4.1 Inert Purge for Tower Overhead with Total Condensation

From the overhead of a tower with total condensation, no vapour product is withdrawn.
Consequently, any non-condensables that ingress with the feed to the tower will get trapped and
accumulate in the overhead vapour space, especially the vapour space in the condenser and the
reflux drum, which is a dead end for vapour. Non-condensables in this context are referred to as
“inerts”. Inert molecules that don’t condense have no place to go but get held up in the dead
space. (In contrast, in the overhead with partial condensation, the inerts are continuously swept
out of the system and carried out by the steady flow of overhead vapour product.) The presence
of inerts in the vapour space of the condenser lowers the partial pressure of the condensing
vapour, and therefore has a marked detrimental effect on the overhead pressure. For example,
consider the scenario in which condensation occurs steadily at the desired overhead pressure of
1000 kPa (a) and at a condensing temperature of 43 °C on the tube surface. In the absence of
inerts in the vapour space, the total overhead pressure is also the partial pressure of the
condensing vapour. It is the partial pressure not total pressure that drives the condensation
process. However, suppose inerts have accumulated in the condenser vapour space so that they
make up 5 % of the total vapour population at any instant in the condenser vapour space. In the
presence of these accumulated inerts, and at the same total overhead pressure, the partial
pressure of the condensing vapour decreases to 950 kPa (a) (95 % x 1000 kPa (a)). In order to
drive the same condensation at the same condensing temperature of 43 °C, the overhead pressure
has to back up to 1053 kPa (a) (1000 kPa (a) ÷ 95 %) to restore the partial pressure of the
condensing vapour to 1000 kPa (a).

It is that seen that the presence of inerts in the overhead vapour space has a pronounced effect on
the overhead pressure and its controllability. The overhead pressure is the one key operating
parameter that must be held constant at the desired value for the proper functioning of the tower.
Since the volume in the overhead vapour space is small, especially the vapour space in the
condenser which is packed full of tubes, it does not take much inert accumulation there to make
its effect felt.

For tower overhead with total condensation, it is of great importance that the overhead vapour
space be provided with a facility for intermittent inert purge. Figure 1.6 shows an arrangement
for an inert purge. It is a good practice place the purge at a point in the overhead so that the
vapour space in the condenser gets a good sweep during inert purging.

14
1.4.2 Semi-flooded Tube Method: Siphon Effects in Condensate Line

As discussed above, the mechanism of varying the tube area exposed to the condensing vapour
consists of adjusting the liquid level in the condenser. For a horizontal condenser of 1500 mm
diameter, a resolution in the order of 10 mm in the liquid level is required for controllability.
This means control action must be sensitive enough to distinguish a level difference of within 10
mm in the condenser. For a vertical condenser with a tube bundle length (height) of 5000 mm,
controllability can be achieved with a correspondingly coarser resolution in liquid level.

Given this resolution required for controllability, it is important to minimize the siphon effect in
the liquid line between the condenser and the reflux drum, as discussed below. Figure 1.7 shows
the setup of the semi-flooded tube scheme, where condensate produced is drained by gravity
from the condenser to the reflux drum, through the condensate line and the control valve in the
same line; gravity drainage is facilitated by the equalization of pressure in the condenser and the
reflux drum. In this setup it is clear that, if the line from the control valve to the reflux drum is
liquid filled, the pressure drop ∆P CV across the control valve is given by:

∆PCV = hM + h1 + h2

Vapour IN

PC
Pressure equalizing
line

hM : shell side condensate


level manipulated to adjust
hM
tube area exposed to
Alw ays
h1 condensing vapour.
liquid filled

Control
Continuous liquid
valve
column may be
sustained in this
h2 pipe segment
causing undesirable
siphon effect

Liquid OUT
Figure 1.7
Semi-flooded tube method; inherent large undesirable dead
heads (h1 + h2 )

In the force balance above, we assume that the friction loss due to flow in the condensate line is
negligible; that is to say, the condensate line is properly sized. The control valve is sized
accordingly to impose the expected pressure drop given by the above relation. For the purpose
of illustration here, let’s use the following parameters and variable:

15
h 1 = 5,000 mm (fixed)

h 2 = 10,000 mm (fixed)

h M = 300 mm (variable) with resolution of 10 mm.

In normal operation, therefore, the control valve has to absorb (or chew up) the pressure drop of
15,300 mm. It is easy to appreciate that the required resolution of 10 mm is far beyond the
capability of this (rather coarse) control valve. This simply means that pressure control with this
setup cannot possibly be expected to work. Requiring a resolution of 10 mm of a control valve
sized for a normal operating ∆P CV of 15,300 mm is analogous to the impossibility of measuring
a machine part to within the precision of a µm using a meter stick (instead of a micrometer)

In the setup shown in Figure 1.7, the segment of the condensate between the bottom of the
condenser and the control valve is necessarily liquid filled, giving rise to a siphon effect of h 1 =
5,000 mm liquid column. For the purpose this illustration, we have assumed that the segment of
condensate line between the control valve and the reflux drum to be completely liquid filled also;
therefore, the siphon effect due to this segment of the pipe is equal to h 2 = 10,000 mm of liquid
column, i.e. the elevation difference between the control valve and the reflux drum. In practice,
the flow regime in this segment of the pipe may consist of plugs of liquid punctuated by pockets
of vapour rushing up the pipe from the reflux drum; in this case, the siphon effect still far
outstrips the resolution we need for controllability. To make a bad situation worse, the siphon
effect due to this flow regime (plugs of liquid punctuated by vapour pockets) can be erratic, with
the making and breaking of vapour pockets in the line.

1.4.3 Minimizing and Eliminating Siphon Effect in an Existing Condensate Line

In an existing total condenser being retrofitted for the semi-flooded tube control method, we
need to eliminate or minimize the siphon effect in the liquid line. The first segment of the line is
inevitably liquid filled. We can minimize the siphon effect in this segment by placing the control
valve as close as possible to the bottom of the condenser; that is, we minimize the elevation
difference between the condenser and the control valve. The siphon effect in the other segment
of the line can be eliminated completely by a siphon breaker, as shown in Figure 1.8.

With the parameter and variable shown in Figure 1.8, the normal operating ∆P CV is 400 mm (=
h M + h 1 ). A control sized for this operating ∆P CV can entertain the resolution of 10 mm required
for controllability.

16
Vapour IN
PC
Coolant OUT

hM
Pressure equalizing line

Control valve placed as close


Control to condenser as practical
h1 (minimized)
valve
Coolant IN Continuous liquid column
Siphon
breaker cannot be established due
h2 to siphon breaker;
no siphon effect: h2 = 0
hydraulically

hM = 300 mm (normal)
h1 = 100 mm
Liquid OUT
Figure 1.8
Eliminating siphon effects in condensate line to reflux drum

1.4.4 New Design: Minimizing Elevation Difference between Condenser and Reflux Drum

Section 1.4.3 shows the measures we would take to retrofit the overhead pressure control with
the semi-flood tube method - a retrofit situation occasioned by a setup that was originally
intended for the hot vapour bypass control scheme, for example.

Vapour IN
Coolant OUT PC Pressure equalizing line
hM = 300 mm
(operating level) Short vertical run;
continuous liquid column
cannot be sustained;
no siphon effect: h2 = 0
Coolant IN hydraulically
h1 = 100 mm h2
(minimized)
Liquid OUT
Figure 1.9
Horizontal condenser stacked atop reflux drum

In a new design, however, we are given a clean slate to do it right the first time. Unless there are
other overriding considerations, the condenser is place as close as possible to the reflux drum.
This usually means the condenser is stacked on top of the reflux drum. The segment of
condensate line between the control valve and the reflux drum is a short vertical run in which the
flowing liquid is broken up due to its proximity to the vapour space in the reflux drum, and
therefore cannot give rise to any siphon effect, as shown in Figure 1.9. Furthermore a vertical
condenser would be preferable, as shown in Figure 1.10, for the desirable naturally linear

17
relation between the tube area exposed and submergence (liquid level) in the condenser, as well
as a wider span of condensate level to manipulate for control.

Vapour IN

PC Coolant Coolant
IN OUT

Short vertical run; Pressure equalizing


continuous liquid column hM line
cannot be sustained; no
siphon effect: h2 = 0
hydraulically

Tube area exposed = a − bhM


Where: a, b are some constants
Liquid OUT
Figure 1.10
Vertical condenser: tube area exposed to condensing vapour is
proportional to condensate level hM ; also hM has a wider span
than that corresponding to a horizontal condenser

1.4.5 Other Considerations in the Use of the Semi-flooded Tube Method

Pressure Equalizing Line

Integral to the functioning of the semi-flooded tube setup is the equalizing line connecting the
tower top vapour space with the reflux drum, as shown in Figure 1.5. The equalization is to avoid
a trapped or “blocked in” vapour space in the reflux drum. The equalization therefore promotes
smooth drainage of the condenser liquid by gravity to the collecting drum. Since there is no
valve required on the equalizing line, the reflux drum can share the pressure relief valve of the
tower; in this case the equalizing line is sized for pressure relief contingency of the reflux drum;
in most instances, this contingency is fire. This results in considerable savings in capital and
maintenance costs and simplicity in design.

Goose-neck in Siphon Breaker

As discussed above, in a retrofit the siphon effect must be eliminated in the segment of the
condensate line between the control valve and the reflux drum. (In the case of a retrofit, in which
the condenser and the reflux drum are in all likelihood placed at greater elevation difference, a
siphon breaker is required to eliminate any siphon effect.) The point of connection of the siphon
breaker to the line segment and the orientation of the siphon breaker are crucial to its function.

18
The guidelines for the installation of the siphon breaker are shown in Figure 1.11. The goose-
neck height of 600 mm is provided to prevent the ingress of liquid into the siphon breaker that
may be caused by process perturbation.

Routing & elevation of siphon breaker


not critical beyond this point Routing & elevation of siphon breaker
not critical beyond this point
Connect to reflux
drum vapour space Connect to reflux
Liquid
(via equalizing line) drum vapour space
ex condenser
600 mm MIN (via equalizing line)
600 mm MIN
Control (∆ elevation) Siphon breaker
(∆ elevation)
valve (goose-neck)
TOL (top of line)
Siphon breaker
(goose-neck)
Liquid Liquid to reflux drum
Minimize Control
ex condenser (open channel flow )
Liquid to reflux drum valve
(open channel flow )

(a) Control vavle in vercal run (b) Control valve in horizontal run

Figure 1.11
Siphon breaker: goose-neck to ensure breaking of siphon effect in liquid line

19
Chapter 2:

STEAM AS A HEATING MEDIUM

2.1 Introduction

In petroleum refineries and other hydrocarbon processing and chemical process plants,
steam condensation is used in large quantities and at numerous locations as a heating
medium. In this application the heat duty of an individual heat exchanger may be as large
as 100 GJ/h.

Steam is usually transmitted and delivered to the users in the superheated state. In order
to use steam condensing as heating medium in a heat exchanger, the steam is first
desuperheated to saturation by spraying boiler feed water (BFW) directly into it. The
saturated steam so produced is then introduced to the heat exchanger as the heating
medium. This ex situ desuperheating to saturation by spraying it with BFW is necessary
to avoid the need to desuperheat the steam inside the exchanger; desuperheating of steam
using heat transfer surface of an exchanger is a highly inefficient process requiring an
inordinately large exchanger tube area. The ex situ desuperheating means that only
surface area for condensing need be provided in a heat exchanger in which only the heat
obtained from steam condensation is transacted. This greatly reduces the surface area,
hence the size of the heat exchanger, required for a given duty. As a rough illustration of
the effect of ex situ desuperheating, consider steam supplied at 1000 kPa (g) and at 185
°C. This steam is typical of an MP (medium pressure) steam available in a refinery or
heavy oil upgrading complex. At 1000 kPa (g) the saturation temperature of steam is 180
°C; the MP steam so supplied therefore comes with 5 °C of superheat. If this steam were
fed directly in the exchanger without any ex situ desuperheating to saturation, the surface
area required of the exchanger would be about 20 to 30 % greater than that required if the
steam is first desuperheated ex situ before entering the exchanger. The extra 20 to 30 %
of tube area is required to desuperheat the steam in situ by 5 °C to bring it to saturation,
in order for the subsequent condensation to take place.

In this chapter we discuss the process control for desuperheating steam ex situ to
saturation, as well as the control of the heat flow rate in a heat exchanger using steam
condensation as heating medium.

2. 2 Ex Situ Desuperheating to Saturation

To bring superheated steam to its saturation temperature by mixing it with a spray of


BFW is a simple process. The goal of control for desuperheating to saturation is to spray
enough water into the steam so that it becomes saturated without excessive free water
(moisture) present in the product steam. The free water (moisture) content of a saturated
steam is known as its “quality”, which is stated in %. A 100 % quality steam is saturated
steam with zero moisture content; while a 0 % quality steam is just saturated water. In
most applications where saturated steam is used it is desirable to have it at as high a

20
quality as practicable. This is certainly the case when saturated steam is used as a heating
medium in a heat exchanger, where it condenses to yield its latent heat as duty. Low
quality steam (i.e. unduly wet steam) is undesirable in more ways than one. On a unit
mass or volume basis, wet steam has lower enthalpy than dry saturated steam. To satisfy
a given duty of an exchanger, a greater quantity of wet steam is required. The excess
water in the wet steam just takes a free ride through the exchanger, contributing nothing
to the heat transaction. The free-riding water wets the exchanger tube surface; this
adversely affects condensation by depriving direct contact between the condensing steam
and tube surface. The free-riding water burdens or even overwhelms the control valve for
condensate disposal. If the wetness is excessive (in a 70 % quality steam, say), the control
valve sized on the basis 100 % quality steam is likely to be overwhelmed; in other words,
in an attempt to dispose of the excessive condensate, the control is driven to 100 % open,
thereby losing control and resulting in condensate back-up in the exchanger and the
attendant operation disruption. Also, in a line supplying excessively wet steam, water
tends to drop out resulting in hammering.

2.2.1 Universal Control Scheme for Making Saturated Steam

Shown in Figure 2.1 is the scheme offered by most engineering firms and technology
vendors as the control for producing saturated steam from superheated steam. The
scheme is almost universal. (The present author has not encountered a different scheme
offered for such control in his 37-year career as a process engineer.) It looks innocuous
and innocent enough; but it does not work, a fact known to all process operators but
categorically ignored by process designers!

: Desuperheater 180 1
(steam-w ater mixer) Saturation line
BFW : Boiler feedw ater 2 (constant temperature)
(demineralized) (w et steam)
3 5
4
Temperature, C

154
o

Product 6
TC
Superheated steam
Saturated steam: Saturated w ater
steam
1 1 100 % quality steam 0 % quality steam
2 (0 % moisture) (100 % moisture)
3 4 5
BFW Water Injection rate
6
(a) Control set-up (b) Product steam temperature versus w ater injection rate:
once saturation is reached at 3 , further injection w ater
causes no temperature change until steam quality
Figure 2.1: becomes zero at 5 .

Single-point temperature control - a fundamentally flawed control scheme for making


saturated steam from superheated steam

It is easy to see why an apparently “straight forward” scheme like this does not work.
Shown in Figure 2.1 is the state (temperature and quality) of the product steam versus the

21
rate of water injection. For this illustration, let’s use the LP (low pressure) steam, which
is supplied as superheated steam (428 kPa (g), 165 °C). The temperature of the
superheated steam decreases, but remains superheated (State 2, for example), as the
sprayed water is introduced into it. As the water injection rate increases, a point is
reached when the product steam becomes saturated at 100 % quality (represented by
State 3). The saturation temperature is 154 °C (at 428 kPa (g), the pressure at which our
LP steam is supplied). Because the steam/water system is single-component, addition of
water beyond this point will not lower the temperature of the product temperature but
only increases its moisture content, as represented by State 4. The temperature of the
product steam remains constant at 154 °C until State 5 (0 % quality) is reached, at which
state all the heat (sensible and latent) of the inlet superheated steam is used in heating up
the spray water to its saturation temperature at 428 kPa (g).

2.2.2 Failure of the Universal Control Scheme in Operation

The universal “single-point” temperature control (Figure 2.1) can fail in two ways
depending on the inherent error of temperature measurement and how close the set point
is to the true saturation temperature. Let’s assume an inherent measurement error of ±2
°C for the temperature; this is reasonable for a temperature sensing device intended for
the temperature range of 100 – 200 °C in our illustration above; this error is also within
the resolution power of modern digital controller action.

If the error is -2 °C (in other words, temperature measured and fed to controller is 2 °C
below the true process temperature), the product steam will be insufficiently
desuperheated; that is, it will still be superheated by 2 °C. The heat exchanger using
steam condensing is sized with no provision for desuperheating on its tube surface.
Desuperheating (to saturation) by 2 °C can cause the heat exchanger to lose about10 to 15
% of its design duty, duty which is meant to be fulfilled by condensing alone.
Desuperheating uses up (or hijacks) an inordinate tube surface and contributes little to the
duty. Consider the LP steam in our example: the latent heat of condensation is 2100.7
kJ/kg; the sensible heat of desuperheating is 1.9 kJ/kg °C. By inspecting these properties
of steam, we can say as a good approximation that, when 15 % of the tube area is
hijacked by in situ desuperheating, the exchanger effectively loses 15 % of its capacity.
Usually this substantial loss of capacity produces noticeable undesired effects on the
process that uses this heat exchanger, for example deterioration of product quality due to
insufficient reboiler duty in a fractionation tower. It is, however, quite another matter
whether or not the cause of this deteriorated process performance is properly identified as
insufficiently desuperheated steam. More often than not, a host of other causes is
suggested, such as malfunction of tray hydraulics, exchanger plugging, exchanger
fouling, etc; after all, who would suspect the good, hot clean steam? This usually results
in disruptive and costly investigations, in the fashion of a wild goose chase, which lead to
no solutions to the problem, while operation is being severely hampered. All this for a
simple and apparently innocuous control scheme for desuperheating steam!

22
If, on the other hand, the error is +2 °C, excessively wet steam will be produced and
delivered to the users with the attending consequences, which include hammering in the
steam supply piping to the exchanger and severe shortfall in duty, as discussed above. As
it can be appreciated easily, with an error of +2 °C, the control valve in the BFW line will
be driven to full open position in a vain attempt to satisfy the erroneous product steam
temperature. (The control valve in the BFW line is relatively small since it is sized only
for the flow rate of BFW to desuperheat the steam.)

Quite apart from inherent measurement errors just mentioned, the saturation temperature
is estimated from a pressure. Depending on who did the estimating, this pressure may be
the steam supply pressure measured near the desuperheater; it may be the nominal
pressure of the steam; or it may be a pressure considered to be “typical” of the steam in
question. From this pressure the saturation temperature is obtained from the steam table,
a chart or some correlation; the saturation temperature so obtained is used as set point for
the controller. One can surmise that the set point may be close to but not exactly equal to
the true saturation temperature at the actual operating pressure of the steam in the
desuperheater. If the set point turns out to be higher than the true saturation temperature,
the product steam will still be superheated; if it is lower than the true saturation
temperature, the product steam will be excessively wet.

2.2.3 Proper Control Scheme for Making Saturated Steam

Once it is recognized that it is futile to use the product steam temperature as the control
target to produce saturated steam, a workable scheme can be easily devised, as shown in
Figure 2.2.
Required only if Optional
pressure letdow n BFW
is required
TT

SP FC
PC CALC

PT TT FT
Saturated
steam

Desuperheater
Steam supply
header (superheated)

Figure 2.2:
Making saturated steam: a scheme that works to maximize
product steam quality

Here the flow of the superheated steam entering the desuperheater is measured; also
measured are its pressure and temperature. All these measurements are fed to a
calculation routine CALC in the digital control system. In CALC, the measured flow is

23
corrected to give the actual mass flow rate of the superheated steam; also calculated is the
mass flow rate of water required to saturate the incoming superheated steam to 100 %
quality (zero moisture). The required water rate so calculated is used as set point for the
flow controller of the water spray.

It is prudent to err on the good side by using 1.03 times the calculated required water rate
as the set point for the water flow controller. It is by far preferable to have a product
steam that is saturated but with a small moisture content than one that is insufficiently
desuperheated; as we have discussed above, 1 °C of superheat in the steam as heating
medium causes a significant reduction in the operating capacity (duty) of the exchanger
designed for steam condensation. This factor of 1.03 (also known as bias) should also
compensate for inherent errors in measurement and in steam/water property data used in
the calculation; and the bias can and should be adjusted in each specific case based on the
knowledge of actual instrument accuracy gained from operating experience to ensure
high quality product steam.

In Figure 2.2, the measured (real-time) temperature of the BFW (boiler feedwater) is
shown as an optional input to the calculation routine; however, in many installations, the
measurement of this temperature is not provided; in this case a known typical
temperature of BFW, which usually remains nearly constant, can be used in the
calculation without introducing significant error. Also shown in the scheme is a pressure
control loop to step down from the pressure of the steam supply header to the desired
pressure (hence desired temperature) of the product saturated steam, should this step-
down be necessary.

The calculation routine is shown in Appendix I: Set Point for Desuperheating Water
Control. The steam/water property data are available in the data base of a modern DCS,
which comes with software that can be programmed to access the data base and to
perform calculations such as this.

2.3 Control of Heat Input with Steam Condensing as Heating Medium

For the control of heat input, many designers’ technique of choice is to throttle the steam
flow at the inlet to the exchanger. Also, in many books and papers, this technique of
choice is shown or implied. When authors are illustrating process schemes that include a
heat exchanger in condensing steam service, they frequently show a flow control valve on
inlet steam, e.g. Friedman (7). Unfortunately, an allegedly symbolic representation of this
sort has inspired many designers to throttle inlet steam when choosing a technique for
control of steam flow in a condensing service. The shortcomings of this approach are
discussed below, followed by a discussion of more robust control schemes.

24
2.3.1 Throttling the Steam Inlet: the Almost Universal Technique

With the schemes shown in Figure 2.3, the pressure of the steam inside in the exchanger
is indeterminate, depending as it does on the extent of throttling of the steam inlet. This
pressure will be erratic if the process heat load varies. Furthermore, at low heat demand
by the process when the steam inlet valve is throttled down, this pressure may at times
become too low to deliver the condensate into the condensate collection header, because
the steam inlet valve is throttled down to a small opening. When this occurs the steam-
side of the exchanger is flooded with condensate back-up, and the heat transfer surface
becomes severely curtailed. The process temperature target is missed, setting up a cycle.
The temperature controller opens up the steam valve, raising the steam pressure and
pushing the condensate into the condensate collection header. Thereafter the steam inlet
throttles back down, condensate accumulates, and the cycle repeats. This erratic heat
transfer and pressure surge that accompanies it will result in process upset. For example,
if the exchanger is a reboiler of a fractionation tower, this erratic behavior will result in
failed fractionation brought about by erratic tray loadings and pressure surges in the
tower.

TC
TC SP
FC

Process OUT Process OUT


Saturated Saturated
steam IN steam IN
LC LC

Condensate Condensate
level level

Condensate OUT Process IN Condensate OUT


Process IN

(a) Simple TC loop (b) TC-FC cascade

TC
From process
SP (somew here)
FC

Process OUT
Saturated
steam IN
LC
Condensate
level

Process IN Condensate OUT

(c ) TC senses temperature from somew here in the process

Figure 2.3:
Heat input to process is controlled by throttling inlet steam yielding steam of
indeterminate pressure inside the exchangers

25
In the schemes represented by Figures 2.3 (a) and (b), the temperature of the process fluid
leaving the exchanger is used as the control target. In (a) the temperature controller (TC)
output manipulates the steam valve directly; whereas in (b) the TC output is used as set
point for the flow controller (FC); it is the output of this FC that manipulates the steam
valve. In Figure 2.3 (c), the temperature of some relevant point in the process
(temperature of tray in a fractionation tower, for example), is used the control target.

The schemes in Figure 2.3 rely on two mechanisms, of about equal prominence and at
work simultaneously, to adjust the heat transfer rate:

• Reducing the effective heat transfer area

Throttling of the steam at the inlet reduces the pressure and this in turn makes the
steam entering the condenser superheated. The resulting requirement for
desuperheating this steam inside the exchanger reduces the tube surface available
for steam condensing. The tube area “hijacked” by steam desuperheating is
effectively lost area, because of the sensible heat transaction associated with
desuperheating, which is insignificant (by 2- 3 orders of magnitude, as discussed
above) compared to the latent heat of condensation; and

• Lowering the driving temperature

At the lower throttled down pressure, steam has to condense at a lower


temperature thus reducing the temperature difference that drives the heat transfer
across to the process side.

Quite apart from disadvantage of operating the steam side of the exchanger at an erratic
and indeterminate pressure that these schemes entail, the reliance on more than one
mechanism to achieve a control objective is undesirable, as this usually leads to highly
non-linear and sporadic response.

2.3.2 Heat Input Control by Manipulating Heat Transfer Area

A key characteristic of a robust control is that only one single prominent mechanism is
exploited. There may be minor unintended mechanisms at play, but they have
insignificant effects on the robustness and response to control action. In the following we
discuss a method of heat input control that avoids throttling steam inlet, thus ensuring a
constant temperature of condensation and a constant pressure to drive condensate to the
collection header. This method employs only a single mechanism, namely a variable heat
transfer area.

The heat transfer rate (here manifested in the process OUT temperature) is controlled by
varying the tube surface exposed using the TC-LC cascade scheme, as shown in Figure
2.4. Condensation occurs at constant temperature since the steam side pressure, riding
freely on the steam supply header, is constant. This constant pressure prevailing in the

26
steam-side of the exchanger is always high enough to deliver the condensate into the
condensate collection system.

Process OUT

Saturated steam IN
TC
Floats freely on constant steam
supply header pressure; no
SP
control valve; no throttling
LC
Condensate
level

Process IN Condensate OUT

Figure 2.4:
Steam condensing in shell side; process outlet temperature
responds to varying heat input

The heat transfer rate is essentially directly proportional to the area exposed to steam
condensing. While in the submerged portion some subcooling of the condensate takes
place, the small sensible heat transaction associated with this subcooling is a secondary
mechanism and is insignificant compared to the large latent heat transacted by
condensation occurring on the exposed tube area. This underscores the fact that the
robustness of control is achieved by relying on one single predominant mechanism – in
this case, the nearly linear relation between the area exposed and the heat transfer rate.
For a vertical tube bundle the area exposed itself varies linearly with the condensate
level, thus conferring to the control dynamics a highly desirable linear response.

The TC, shown at the process side outlet in Figure 2.4, may have its process temperature
sensed at some other point in the process (a fractionation tray, for example) where the
temperature responds meaningfully to varying heat input in the exchanger. The scheme
shown in Figure 2.4 is therefore suitable for applications where the process side
temperature changes sufficiently with the rate of heat input. For example, this is the case
for heating up a fluid without phase change, or partially boiling a liquid with a
sufficiently wide boiling range (such as in the reboiler of a multicomponent fractionation
column).

It cannot be over-emphasized that in the above and the subsequent schemes, it is


important to prevent steam blow-out into the condensate collection system. We shall
discuss in greater details the technique to implement this measure in a separate section.

The schemes in Figure 2.5 are a variation of the same concept expounded above in Figure
2.4. The main consideration here is that the process fluid has a narrow boiling range; in
other words it boils at a constant temperature, or over a very narrow range of
temperatures. Hence this temperature does not respond sufficiently, or at all, to varying
heat input. Here the heat input rate (steam flow rate) is controlled by varying the tube
surface exposed to condensing using FC-LC cascade (instead of the TC-LC cascade in

27
Figure 2.4). The set point for the FC may be determined, for example, by feed rate or
feed temperature to a process unit, or both; or some other process variable the target of
which is to be met by varying the heat input. These schemes are, for example, suitable for
the heat input control for a reboiler of a sour water stripper. In these applications, the
process liquid boils in the reboiler at a nearly constant temperature, regardless of heat
input rate.

Note that in Figure 2.5 (a) we have the process fluid in the shell side and steam
condensing in the tube side, in a kettle-type heat exchanger. The same technique applies
to other arrangements, such as a vertical exchanger with process fluid boiling in the tube
side, as shown in Figure 2.5 (b).

Saturated steam IN
Process OUT Floats freely on constant supply
(Vapour) header pressure; no control
valve; no throttling
Process
Saturated steam IN OUT (Liquid
Floats freely on + vapour)
LC constant supply header LC

SP pressure; no control
Process IN FC valve; no throttling SP
FC
Process OUT
(Liquid) Condensate level Condensate Condensate level
OUT Condensate
Process IN
OUT
(a) Horizontal tube bundle: steam condensing (b) Vertical tube bundle steam condensing
in tube side in shell side
Figure 2.5:
Process fluid boiling at constant or nearly constant temperature independently
of heat input; heat input rate control is achieved by varying the tube area
exposed to the condensing steam

Unlike the vertical tube bundle shown in Figure 2.4, the tube area exposed for a
horizontal tube bundle, as shown in Figure 2.5 (a), is not related linearly to the
condensate level inside the tube bundle. However, to obtain a desirable linear response
the level controller signals can be easily linearized in the control algorithm using the
geometric relation between the tube area exposed and condensate level inside the tube
bundle. Similarly, for an exchanger condensing in the shell side a with horizontal tube
bundle, the condensate level in the shell does not have a linear relation with the tube area
exposed to condensing. Here, too, the LC output can be linearized using the geometric
relation between the liquid level and the tube area exposed in the shell side.

28
2.3.3 Measure to Prevent Steam Blow-out

Steam blow-out into the condensate collection system can inflict severe damage to
equipment and strike fear into those who experience it. All control for steam
condensation must include measures to prevent this horrifying event from happening

Figure 2.6 shows the same basic scheme as in Figure 2.4 but with HSS (high signal
selector) and HS (“hand switch”) added to prevent steam blow-out into the condensate
collection system. The lowest allowable condensate level, say 5 %, is set by the operator
using HS. (HS, which stands for “hand switch”, is a hangover from the olden days when
this function was executed with a physical hardware called a “switch” located on a
control board in the control room or out in the field. The operator would set the parameter
by turning a dial or moving a slider on HS. Nowadays, the function of HS and those of
other functional units, such as TC, FC, LC and HSS, in the control scheme are all
performed digitally by programmable process control software.) The output signals of TC
and HS compete to become the set point of LC. The higher of the two is selected by HSS
to pass on to LC as its set point.

Process OUT
Saturated steam IN
TC
Floats freeely on constant header
pressure. No control valve; no
throttling SP
Condensate
LC HSS HS

level

Condensate OUT
Process IN
TC and HS outputs = % level set point for LC
HS output = minimum allow able condensate level
l l
Figure 2.6:
Control scheme to prevent steam blow-out; the HSS guarantees a
minimum allowable condensate level in the exchanger

In the alternative shown in Figure 2.7, the minimum allowable condensate level is given
by the operator as the set point for LC. The outputs (% valve open) of LC and TC
compete to reach the condensate level control valve. The LSS (low signal selector) allows
the lower of the two outputs to pass through to modulate the control valve position. In
other words, the controller (TC or LC) that calls for smaller valve opening has its wish
come true.

Both methods have been used successfully. It is interesting that most designers adopt the
scheme shown in Figure 2.6, which they feel is more “intuitive”.

29
Process OUT
Saturated steam IN
TC
Floats freely on constant header
pressure. No control valve;
no throttling
Condensate LC LSS
level

Condensate OUT
Process IN
TC and LC ouputs = % valve open
LC set point = minimum allow able condensate level
Figure 2.7:
Alternative method to prevent steam blow-out; LSS guarantees
a minimum allowable condensate level in the exchanger

For the sake of completeness, we show in the Figures 2.8 to 2.11 the cases in which
process side temperature is not suitable as a control target for heat input control; that is,
the process fluid boils at nearly constant temperature regardless of heat input rate. The
narratives for these schemes in Figures 2.8 to 2.11 are the same as those in Figures 2.6
and 2.7, with TC replaced by FC.
Steam IN
Process OUT Floats freely on constant header pressure
(vapour) No control valve; no throttling

HSS HS
SP
LC

Process
OUT (liquid) FC
Process IN

Process fluid boiling at constant Condensate OUT


or nearly constant tem perature
Figure 2.8: FC and HS outputs = % level

Measure to prevent steam blow-out for process temperature not responding to heat input rate;
horizontal tube bundle: HSS guarantees a minimum condensate level in the exchanger

Saturated steam IN
Process OUT
Floats freely on constant header pressure
(vapour)
No control valve; no throttling

LSS
LC

Process
OUT (liquid) FC
Process IN

Process fluid boiling at constant Condensate OUT


or nearly constant tem perature
FC and LC output = % valve open

Figure 2.9:
Alternative to scheme in Figure 2.8

30
Process OUT (Liquid + vapour)
Saturated steam IN
Floats freely on constant header HSS HS
pressure. No control valve; no SP
throttling LC

Condensate
level FC

Process IN Condensate OUT


FC and HS outputs = % level
Process fluid boiling at constant or
nearly constant tem perature

Figure 2.10:
Measure to prevent steam blow-out for process temperature not
amenable to heat imput control, vertical tube bundle; HSS
guarantees a minimum condensate level in the exchanger

Process OUT (Liquid + vapour)


Saturated steam IN
Floats freely on constant header
pressure. No control valve; no LSS

throttling LC

Condensate
level FC

Condensate OUT
Process IN
FC and LC output = % valve open
Process fluid boiling at constant or
nearly constant tem perature

Figure 2.11:
Alternative to scheme in Figure 2.10

31
Chapter 3:

SEPARATORS
(For gas and liquid separation by gravity)

3.1 Introduction

Separators are vessels which provide residence time for gas and liquid to separate and
provide means of directing the flows inside the vessel and collecting the separated
product streams. The liquids may contain a small incidental amount of solids, such as
sand. The separation is basically driven by gravity. There may be devices placed inside
the separator vessel to enhance coalescence thereby shortening the residence time and
consequently the size of the separator vessel. These devices fall in general in two
categories: “passive” and “active”. The passive type comes in the forms of vanes, plates,
grids or meshes, to create a tortuous path for the droplet-bearing fluid to flow. This
tortuous path increases the chance of the entrained liquid droplets colliding and
coalescing. In the active type, an electric field is maintained across the tortuous path. An
arrangement to provide this electric field in the vessel is known in the trade as the “grid”.
The electric field induces each droplet to become polarized. The polarized droplets attract
one another resulting in coalescence. This greatly augments the coalescence brought
about by collision alone. An active coalescing device can shorten the residence time
required to minutes in some applications where it would take hours or even days if
droplet collision alone is the operative mechanism. The grid is used for crude oil
“desalting” and for heavy oil-water separation in vessels known as oil treaters. Desalting
here means the removal saline water or sea water entrained in crude oil.

These coalescing devices, active or passive, simply enhance the fundamental gravity
driven separation. They are used when justified. They shorten the residence time but do
not change the underlying design principle of the separator; hence in this chapter we do
not present details of coalescing devices. We look at a range of applications of the
separator, which can be covered as follows:

• Two phase separation (gas and one liquid phase)

• Two phase separation (two liquid phases; gas is absent or incidental)

• Three phase separation (gas and two liquid phases)

The arrangement of internals (baffles, weirs, boxes and compartments) in the vessel
varies greatly in this range of applications; so does the arrangement for the introduction
of feed and the withdrawal of products. These variations depend on:

• Whether the separator vessel is vertical or horizontal;

• The relative rates of the fluids (gas, first liquid and second liquid) to be separated.

32
In this chapter, the principle and calculation procedure for sizing the separator and for the
design of its internals will be outlined. The governing processes and equations will be
presented. Given these, the reader can easily set up his own calculations in a spreadsheet
or other programmable software. The primary objective here is to highlight and discuss
the salient points often missed in the commercial design of separators offered by
engineering houses and technology vendors.

3.2 Horizontal versus Vertical Separator

The separator in the context of this chapter is a cylindrical vessel which can be installed
in a vertical or horizontal orientation. From the point of view of separation alone, a
horizontal separator always provides greater residence time than a vertical one, for a
given vessel volume, due to the more effective flow paths of the separated oil and water
drops afforded by the horizontal vessel as discussed in Section 3.2.1 below. However,
other considerations may lead to the choice of a vertical separator:

1. Horizontal separators are not as good as vertical separators in handling solid-


bearing liquids. Periodic removal of accumulated solids – an operation known as
solids dumping - is more elaborate and usually less effective for a horizontal
separator than it is for a vertical separator.

In a horizontal oil treater, "de-sand" nozzles have to be placed at close and regular
intervals due to the angle of repose of the solid deposit (about 45 degree for sand).
Success is dubious with de-sand jets used to loosen sand while dumping, whereas
it is easier to loosen and fluidize sand (or other solids) collected at the bottom of a
vertical treater with water jets during solid dumping.

2. A horizontal separator requires more foot print (plot area) - an important


consideration for off-shore installations or skid-mounted installations where foot
print is at a premium.

There are, however, other considerations against the use of vertical separators, quite apart
from their inherent lower efficiency in providing residence time (see Section 3.2.1
below), compared to horizontal separators:

1. In the skid mounted construction, all pieces of process equipment and ancillary
(instrument and electrical, etc) are mounted and tested in the shop so that the
whole assembly on the skid is ready to function when it is transported and set in
the field. However, due to its height a vertical separator often has to be shipped
loose and separated from other skid mounted equipment and erected in the field.

2. Access to the top or high elevations of a vertical separator (for operation and
maintenance, etc) in the field need to be considered: ladder, cage, guard rails,
platform.

33
3. Many oil and gas production facilities are housed in a heated enclosure. A vertical
vessel often has its top part sticking out of the roof of the enclosed space
(building), creating additional requirements for winterization and access.

3.2.1 Inherently Higher Efficiency of Horizontal Separator over Vertical Separator

It is easy to appreciate why a horizontal separator is inherently more efficient than a


vertical one in providing residence time. As shown in Figure 3.1, in a vertical separator
liquid drops in the gas phase have to fall head-on against an upward current of gas. Oil
drops that coalesce in the water phase have to rise against a descending current of the
water phase. The water drops in the oil phase have to fall against an upward current of the
continuous oil phase. In contrast, in a horizontal separator, these drops do not fall or rise
directly opposite to the directions of the currents of the phases, but perpendicular to them.
This inherently higher efficiency transfers to longer residence time for a given volume;
this can be verified using the design and sizing procedures for separators, such as those
presented in this chapter.

Gas Liquid drop falling in gas phase

w ater drop falling in oil phase


Gas phase Feed
Oil drop rising in w ater phase

Flow direction of gas, oil, w ater in their


respective phase
Oil
Oil Feed Gas
phase

Gas phase
Water
phase Oil phase

Water phase

Water Oil
Water
(a) Vertical separator: separated droplets rise (b) Horizontal separator: separated droplets rise
or fall counter current to the motion of the or fall cross current to the motion of the phases
phases

Figure 3.1:
The inherently higher efficiency of the horizontal separator over the vertical by virtue
of the more favorable flow directions of the gas, oil and water phases with respect to
the separated droplets falling or rising in them

34
3.3 Horizontal Separators – Category of Applications

General Principle

The requirements of the general principles governing the sizing and design of a horizontal
separator are:

1. The inlet must discharge into the vapour space so that the bulk of the gas breaks
out directly into the vapour space above the residence liquid surface, to minimize
agitation of the liquid body in the vessel;

2. The inlet must be introduced at one end of the separator, as close to the end head
as practical, to avail of the vessel volume for residence time for the separation of
the two liquid phases and the disengagement of any remnant dissolved gas; and

3. The separated fluids, namely vapour and one or two liquids, are to be withdrawn
from the other end of the vessel, as close to the end head as practical; this in
conjunction with the location of the inlet makes the most use of the vessel volume
for residence time.

Vapour Disengagement

The sketch of one scheme for vapour disengagement of the inlet fluid is shown in Figure
3.2.

(1)
Min. Inlet nozzle
centerline
Impingement
(2) 25
plate Liquid
D/12
(4) Liquid surface
150
surface
min.
Liquid-liquid
Splash Splash guard to have slots across
Interface
(3)
baffle 100 its w idth (only tw o show n)
(5)

(1) Minimum distance subject to fabrication constraint.


(2) Impingement plate diameter to be 2 x inlet diameter minimum.
(3) Splash guard to extend w all to w all
(4) D/12 or 150 mm, w hichever is greater
(5) Slots to prevent accumulation of oil behind splash guard

Figure 3.2:
Inlet orientation and placement for vapour disengagement in vapour space

A better scheme is to introduce the incoming fluid into the body of the resident liquid
below the liquid surface, after allowing the incoming fluid to disengage the bulk of its
vapour in the vapour space. This applies whether we have a single or two liquids in the

35
feed. This disengagement of vapour in the vapour space is important as this maximizes
quiescence in the liquid body, quiescence that is necessary for oil-water separation – or
any liquid-liquid separation. Maintaining quiescence in the liquid body also minimizes
frothing and foaming, whether we have a single liquid or two liquids. A scheme to
accomplish this is shown in Figure 3.3. This scheme of introducing the feed liquid into
the resident liquid body is superior to the scheme presented in Figure 3.2; with the
launcher to spread the incoming liquid to cover as evenly as possible the flow area of the
resident liquid, closer approach to plug flow (hence closer approach to the theoretical
residence time), as well as a higher degree of quiescence, are achieved than in the case of
splashing the feed liquid onto the resident liquid body as shown in Figure 3.2. Note that
this scheme of introducing the feed liquid also does away with the impingement plate and
the splash guard, replacing them with the gas break-out pan and the launcher. The feed
liquid launcher is presented in some detail in Section 3.4: Feed Launching.

Inlet pipe
(gas-bearing liquid)

Deflector Gas breakout


(1) (3)
cone pan

Liquid surface

Tw isted tape
insert Liquid flow

(2)
Launcher
(1) Right-angled cone diameter = 2 x inlet pipe OD
(2) See Section 3.4 on liquid launcher details.
(3) Pan diameter sized for vapour velocity < 5 ft/s exiting pan.
Figure 3.3:
Arrangement for vapour disengagement for use with
liquid launcher

It should be noted that in many sketches of the horizontal separators presented in this and
other chapters, vapour disengagement of the feed is shown symbolically in the manner of
Figure 3.2, as it is conventional to do so in the literature in general. We strongly
recommend the use the method of vapour disengagement and feed launching shown in
Figure 3.3.

Liquid Withdrawal

In the case of two liquids separating, the liquid withdrawal schemes depend on the
location of the liquid-liquid interface level, which in turn depends on the relative rates of

36
the two liquids in the incoming fluid. The liquid withdrawal schemes fall into four
classes: (a) oil skimming, (b) bulk separation, (c) dewatering and (d) de-oiling. The
vessel internals corresponding to these four classes of operation differ greatly.

(a) Oil Skimming

The separator is designed for oil skimming if a small incidental quantity of liquid
hydrocarbon (oil) is present in the feed. (See Chapter 4: Separator with Oil Skimming
for a fuller treatment of application of the three-phase horizontal separator with oil
skimming.) The aqueous phase is collected and withdrawn by means of an underflow-
overflow weir arrangement. The hydrocarbon liquid is collected by means of an overflow
weir arrangement. A sketch of this scheme is shown in Figure 3.4. The oil-water interface
level is not controlled. The height hL of the hydrocarbon overflow weir relative to that h A
of the aqueous phase overflow weir, i.e. hL − h A , should be designed such that a layer 50
to 100 mm thick of liquid hydrocarbon (oil layer) can form at the quiescent end of the
liquid body in the residence compartment of the separator, based on the predicted
densities of the two liquids at the conditions and throughput of the separator.

Feed PC
Vapour
Oil trough
IN OUT
LC
Liquid surface

Oil layer LC

Interface hL hA

Hydrocarbon Aqueous
OUT phase OUT

Residence compartment (1)


Surge compartment
(1) Compartment volume to be minimized subject to surge requirement

Figure 3.4:
Horizontal three phase separator: Baffle arrangement for oil
skimming; small or incidental quantity of oil in water.

The entrained oil in the inlet is indeterminate but small. This means that the oil phase
residence time (in the oil layer) is also indeterminate but very long; it may be hours or
days. Weathering of the resident oil layer increases the oil density; this is undesirable as
heavier oil tends to break-out, drifts down into the aqueous phase and gets entrained into
the outlet of this phase. So, ideally, the oil layer thickness should be kept as small as
possible to minimize the effect of weathering. But too thin an oil layer could cause the
aqueous phase to overflow into the oil trough due to process perturbations, such as wave
action propelled by a gush of gas in the inlet, or a sudden increase in inlet rate, which
could obliterate the oil layer if it were kept too thin. A certain oil layer thickness needs to
be maintained to buffer process perturbations, as well as to allow for any inherent errors

37
in the estimate of fluid properties and other process parameters which affect the operating
oil layer thickness. Too thick an oil layer, apart from aggravating weathering, constitutes
a waste of residence volume of the vessel. An oil layer of 50 to 100 mm, which runs the
gamut of designer conservatism, is deemed adequate by conventional wisdom.

(b) Bulk separation

The separator is designed for bulk separation when the rates of the two liquids are
comparable. The hydrocarbon phase is collected by means of an overflow weir and
withdrawn on level control in the hydrocarbon surge compartment. The aqueous phase is
withdrawn on interface level control. A sketch of this scheme is shown in Figure 3.5.

Feed Vapour
Hydrocarbon
IN OUT PC
(1)
surge compartment
Liquid surface Liquid
Interface
Anti-vortex surface
INTERFACE
LC LC
(2)
h

Aqueous phase Hydrocarbon


OUT OUT

(a) No water boot

Feed Vapour Hydrocarbon


IN OUT (1)
PC
surge compartment

Liquid surface Liquid surface


Interface
INTERFACE LC
LC
(2)
h

Hydrocarbon
OUT
Aqueous phase
OUT

(b) Water boot required


(1) Compartment volume to be minimized subject to surge requirement.
(2) If h >= 250 mm, no w ater boot is required and an anti-vortex baffle is required
for aqueous phase draw -off. For h < 250 mm, use w ater boot to capture
the aqueous phase for its outlet.

Figure 3.5:
Horizontal three phase separator; baffle arrangement for bulk separation; oil and
water rates are comparable.

If the oil-water interface is close to the bottom of the vessel (< 250 mm), a water boot
should be used to prevent the unintended suction of oil into the water outlet.

38
(c) Dewatering
Anti-jump
Vapour
Feed (1)
baffle OUT
IN PC

Oil surface LC

A Oil-w ater interface


A
INTERFACE
LC Hydrocarbon
OUT
Aqueous phase
OUT
(a) Elevation

Thin w ater layer Water Anti-jump Oil out


or streaks boot (1)
baffle nozzle

(b) View A-A


(1) Anti-jump baffle height to adequately span w ater boot; anti-jump
baffle is to prevent the flow of w ater to the oil outlet nozzle.
Figure 3.6:
Horizontal three phase separator: Baffle arrangement for
dewatering: small or incidental quantity of water in oil.

When a small or incidental quantity of water is present in the feed, the separator is
designed for dewatering. The hydrocarbon is withdrawn downstream of an anti-jump
baffle and is on level control. Water is collected in a boot and withdrawn on interface
level control. A sketch of this scheme is shown in Figure 3.6.

(d) De-oiling

This arrangement for product withdrawals is suitable for applications where the quantity
of oil in the feed is incidental and small, and the quantity gas is also small. Here the flow
of gas, if any, exits the separator with the light liquid phase (oil). Note that there is no
vapour space in the separator; the inlet is introduced into the liquid body. (See Figure
3.7.) If, however, the presence of gas in the feed is substantial, the arrangement for oil
skimming must be used. See (a) Oil Skimming above.

39
Oil (incidental)
+ gas (incidental)

Feed INTERFACE
Oil-w ater
IN LC
interface

(1)
Water

(1) Signal from a level controller at the source of feed to the separator

Figure 3.7:
Horizontal three phase separator; baffle arrangement for
de-oiling with little or no gas

3.4 Feed Launching

Theoretical Residence Time

The separator is a vessel that, by virtue of its volume, provides residence time required
for the process of separation. In every situation there is always a so-called theoretical
residence time. For example, in a three phase separator (gas, oil and water) as shown in
Figure 3.8, the residence volume occupied by water divided by the volumetric flow rate
of water gives the theoretical residence time for water in the vessel; the theoretical
residence times for oil and gas are obtained similarly; hence:

Theoretical residence time of gas = LG AG QG


Theoretical residence time of liquid 1 (oil) = Lo Ao Qo
Theoretical residence time of liquid 2 (water) = LW AW QW

Where: QG , Qo and QW are the volumetric flow rates at the operating conditions of gas, oil
and water, respectively; L G , L o , L W , A G , A o and A W are defined in Figure 3.8.

The theoretical residence time is the maximum achievable. No contrivances, however


clever, can result in a residence time greater than the theoretical. It requires perfect plug
flow of the fluid to occur in each phase, as it traverses its path in the vessel. This in turns
means there must be no back-mixing, no slippage and no channeling along the flow path.
Tall order! The effective, or actual, residence time achieved in a separator is some
fraction of the theoretical. This effective residence time depends on the geometry of the
vessel and the arrangement of its internals (baffles and weirs); and, in particular, it

40
depends on how the inlet fluid is launched and broadcast into the vessel. What can the
designer do to approach as much as possible the theoretical?

Residence compartment Surge


compartment
LG

Gas Feed Gas OUT


A
breakout
Oil overflow w eir

Oil-w ater A
Water OUT Oil OUT
interface
LW
Lo
AG
Gas residence volume = LG A G Ao
Oil residence volume = LO A O AW
View A - A
Water residence volume = LW A W Flow areas of gas, oil and w ater

Figure 3.8:
Residence volumes of gas, oil and water in a three phase separator

Gas Disengagement

Implicit in the requirement for the approach to the theoretical residence time for liquid-
liquid separation is that quiescence must be obtained in the liquid body in the residence
compartment. In the water phase oil droplets coalesce and rise up into the oil phase.
Similarly, in the oil phase, water droplets coalesce and sink down to join the water phase.
This coalescence, whether or not it is enhanced by devices such as coalescing vanes or
electrostatic grids, requires the quiescence of the liquid phases. Agitation, such as caused
by gas disengaging in the liquid body, runs counter to the process of coalescence and
separation, as it tends to re-mix the otherwise separated oil and water phases. To this end,
the gas-bearing feed must be allowed to flash in the gas space to disengage the gas
directly into gas space, without disturbing the residence volume of the liquid. Depending
on the actual gas volumetric rate, the gas space provided need not be exactly half the
volume of the vessel; in other words, the flow area (A G in Figure 3.8) for the gas stream
need not be exactly half the circular cross section.

In many commercial designs, the whole gas-bearing feed is introduced into the body of
the liquid by a device known as distributor or called by some impressive technical-
sounding name. Regardless of the details and “sophistication” of this so-called
distributor, it is clear that the bubbling action of gas disengaging below the liquid surface

41
is not conducive to creating the desired quiescence in the liquid body. One technology
vendor uses this rather bizarre and counter productive way of introducing gas-bearing
feed to the separator as its immutable standard design, and denies the merit of the simple
and direct gas disengagement into the gas space. These designers seem hell-bent on re-
mixing the separated phases by agitating the resident liquid body by bubbling gas into it!

Liquid Launching for Circular Cross Section

After disengaging the gas in the vapour space, the quiescent liquid is then launched into
the resident liquid body. The design of this launching device (launcher) is motivated by
the desire to approach the ideal, i.e. the theoretical residence time. We know that the ideal
can never be reached, but it can always be approached, and the only limit to this approach
is the understanding, or the lack thereof, of the governing principles and the forces at
play.

Feed
Gas breakout Rotational motion
pan of launched liquid

Oil Oil Launcher

A A
(b) View A - A:
Oil-w ater Aggregate of spinning liquid elements is an
Launcher overall rotational motion of the launched
interface
liquid resulting in centrifugal force propelling
Launched Vessel
the liquid to cover the circular cross section
liquid element axis
(typ)
Water
Spin
direction Angular momentum L
(a) Elevation: centrally positioned launcher in a
vertical cylindrical vessel; liquid elements parallel to vessel axis
are launched spinning and headed tow ard
(c) Sense of angular momentum of
the vessel w all
individual launched liquid element
Figure 3.9:
Imparted angular momentum resulting rotational motion of the launched liquid

Take a vertical cylindrical separator (a tank, for example) as shown in Figure 3.9. The
closest approach to the ideal is to launch the quiescent (gas free) liquid into the residence
liquid body in the tank so that the liquid spreads out evenly to cover the whole circular
cross-section of the vessel, without inducing an undue amount of agitation. Given the
circular geometry of the flow area, the simplest way to approach this ideal is to launch the
quiescent liquid at the center of the circle. Each element of the liquid is imparted an
angular momentum as it is being launched; i.e. each liquid element is made to spin as it

42
leaves the launcher. The aggregate of the angular momenta of the individual elements
results in the launched liquid as a whole rotating about the axis of the tank. (See Figure
3.9 (b).) The centrifugal force due to this rotation spreads the launched liquid to cover
evenly the cross section (i.e. flow area) of the tank. We present in Appendix II the
principle of launching with angular momentum; in Appendix III, we present the design
and sizing procedure a launcher based on this principle, as well as the rate of angular
momentum (torque) required to achieve a complete broadcast, i.e. a complete coverage of
the flow area.

Liquid Launching for Truncated Circular Cross Sections

The discussion above pertains to a circular cross section that presents itself in a vertical
cylindrical separator (or a tank). For a horizontal cylindrical three-phase separator, the
flow area for the liquid, i.e. area transverse (or perpendicular) to the flow direction, is not
a full circle but a truncated one. (See Figure 3.10.) A launcher based on the use
centrifugal force (arising from angular momentum) to broadcast the liquid feed to fill the
flow area still applies. In this application, the diameter D used in design and sizing of the
launcher is replaced by the equivalent (or hydraulic) diameter DE , where:
4 × flow area AL
DE =
perimeter of flow area

The launcher should be placed so that its axis is common with the axis through the
centroid of the flow area. For a truncated circular area such the liquid flow area in a
horizontal cylindrical vessel shown in Figure 3.10 , the centroid can calculated by
elementary calculus.

Launcher

Liquid surface
AL hL Flow

(1)
AL = liquid flow area at liquid Launcher Launcher axis
level h L (a truncated circle)
(a) Liquid flow area (b) Launcher elevation and orientation

(1) Launcher axis to be parallel to vessel axis and


passing through centroid of flow area

Figure 3.10:
Orientation of launcher for horizontal cyclindrical vessel

43
3.5 Common commercial feed launchers

As discussed above, it is a simple matter to impart angular momentum to enable a fluid,


launched at the center (or centroid) of the flow area, to spread by centrifugal force to
cover the whole flow area, thus approaching a plug flow. Spreading outward from the
centre toward the circumference of the vessel is what centrifugal force does naturally!
Also stressed above is the importance of complete disengagement of gas and vapour
directly into the vapour space above the resident liquid body in the residence volume, and
the simple ways of accomplishing it. However few commercial designs of distributors or
launchers take these principles into account. We usually end up with a medley of
contraptions that the engineering houses and technology vendors proffer as their standard,
and sometimes “non-negotiable”, fares; quite often these contraptions are touted as
distinct advantages which the client is fortunate to enjoy by using the service of its
promoter.

For a vertical cylindrical vessel, one common attempt to cover as much of the cross
section (flow area) as possible is to use a device in the form of a “wagon wheel”. The
spokes of the wheel are provided with holes through which the liquid, or, more often than
not a mixture of liquid and gas, is launched. (See Figure 3.11.) Most of these launchers
do not provide for gas break-out above the resident liquid body. A common alternative to
the “wagon wheel” is the “fish bone” launcher as shown in Figure 3.12. Both the wagon
wheel and fish bone launchers are known as spreaders in the business, where they are
commonly used in vertical cylindrical tanks for oil-water separation.

Perforated
Perforated "side-bone" (typ)
spoke (typ)

Vessel
w all

Dow ncomer Dow ncomer


pipe pipe

Figure 3.11 Figure 3.12:


Wagon wheel launcher (a.k.a. spreader) Fish bone launcher (a.k.a. spreader)
Total feed
(liquid + gas)
Slots (2)

View A - A
A A

Liquid General flow


surface direction in
vessel

Slot
Figure 3.13:
Slotted pipe launcher (a.k.a. distributor)

44
Others use the slotted pipe as shown in Figure 3.13. In this contraption (promoted by its
proponents as a proprietary “feed distributor”), the feed, without any de-gassing in the
gas space, is introduced into the resident liquid body. The feed is shot out of the two slots
sideway, as well as vertically downward, into the liquid body. The parts of the slots
above the liquid surface are intended as an attempt to de-gas the feed. As this is a
standard offering regardless of the gas content of the feed, then, depending on the gas
content in the feed, there may still be gas bubbling into the liquid body creating the “egg-
beater” effect, which is detrimental to the quiescence required for separation. In any
event, with gas breaking out so close to and below the liquid surface, it is not conducive
to establishing quiescence in the resident liquid body; it also fails to spread the liquid to
cover the cross section of the separator.

However, one interesting design does attempt to use angular momentum and the
centrifugal force engendered by it in a so-called tangential launch (or tangential feed
technique, as its designers promote it). This launcher is intended for a vertical cylindrical
vessel. In this technical-sounding contraption as shown in Figure 3.14, the liquid or, more
likely a mixture of liquid and gas, is launched as a single jet at a point as close as possible
to the circumference (inner wall) of the vessel. The launch nozzle is trained to give the jet
a tangential velocity at the launch point – hence tangential in the name. Unfortunately,
the centrifugal force engendered by this tremendous angular momentum tends to press
the launched liquid against the circumferential wall of the vessel and cause it to move
down as an annulus in the liquid body. This tends to create a dead zone of liquid in the
core of the vessel, thus reducing the effective residence time for oil-water separation. For
want of a better name, one could perhaps describe this undesirable situation as “annular
channeling by design”. Yes, the notion of centrifugal force is incorporated in the design,
but the centrifugal force is administered at the wrong point!

Vessel w all
(circumference)

Feed Top view - vertical tank

Tangential Jet launched tangential to vessel


launcher circumference and in the plane of
(jet nozzle) vessel cross section
Figure 3.14:
Tangentail launcher

Another contraption uses a maze of baffles (in a wet oil dewatering tanks, for example) in
the hope of forcing the launched liquid to cover the whole cross section of the tank.

45
(See Figure 3.15.) The feed is launched as a horizontal jet into the residence volume. This
is hardly conducive to approaching plug flow; instead, it results in severe channeling and
a waste of volume provided by the tank. The effective residence turns out to be a small
fraction of the theoretical, because the fluid channels at high velocity through the
meandering path.

Oil trough Baffles

OIL
OUT
Launcher
A A (jet nozzle)
Water
OUT
Launcher nozzle
Feed (jet nozzle)
(a) Plane view (b) View A - A : Elevation

Figure 3.15:
Maze of baffles: straight horizontal jet from launcher results in severe
channeling and waste of residence volume

Some advocates of this “maze of baffles” design claim that it provides longer residence
time than given by the theoretical: θ = V Q , where θ is the theoretical residence
time, V the residence volume (volume occupied by the residence liquid body in the
vessel) and Q the liquid rate in the feed. They consider this a self-evident fact, because the
liquid path, hence liquid residence time, is made longer by forcing it to meander through
the maze. We know this claim is not valid because the residence time given by θ = V Q is
the maximum possible and, in practice, it can be only approached, but never reached
(much less surpassed), no matter how clever a contraption is employed. Furthermore, the
only hope of approaching this theoretical maximum is to launch the fluid in such a way
that it tends to assume a plug flow pattern. Launching the feed as a jet, as this technique
calls for, is antithetical to promoting a plug flow across the flow area. The trajectory of
the jet, which has a high velocity, constitutes severe channeling leaving much of the
residence liquid body as wasted dead zone.

It should also be noted that in this medley of contraptions, proper gas break-out in the
vapour space is usually ignored when it is required. In addition to the loss of residence
time due to improper launching, separation is further hampered by the destruction of
quiescence in the resident liquid body by the gas bubbling through it.

46
3.6 Design of Separators

In the flowing subsections we present the design outline for three-phase and two-phase
separators, both horizontal and vertical. We exclude the horizontal three-phase separator
with oil skimming, which is treated exclusively in Chapter 4.

3.6.1 Horizontal Three-Phase Separator for Bulk Separation

Design Considerations

The vessel dimension (length x diameter) should be such that:

(1) There is enough residence time in the gas phase for the falling liquid droplets to
reach the liquid surface;

(2) There is enough liquid volume to meet the desired water residence time;

(3) There is enough liquid volume to meet the desired oil residence time;

(4) The resultant oil layer thickness is smaller than the maximum thickness allowable.
The limit to oil layer thickness arises because the water droplets entrained in the
oil layer must have enough time to sink into the water phase below.

The water layer thickness, however, does not pose a constraint in this application
because the rate of ascent of oil droplets in the water layer into the oil phase
above is far greater than the rate of descent of water droplets through the oil phase
above. This is so because, an object moving in the oil phase experiences far
greater viscous drag than it does moving in the water phase; and

LTT
LG Gas/vapour
Inlet (1)
LL
Vt
PC

hL
D INTERFACE
LC LC

Controlled Water Oil


oil- w ater interface
LS
Oil compartment
(1) LL also marks liquid residence compartment
Figure 3.16:
Bulk separation of gas, oil and water

47
(5) There is enough volume in the oil compartment for oil inventory surge to facilitate
control. (See Figure 3.16.)

Procedure

The process parameters for which the separator is designed are:

Gas volumetric flow rate at conditions: QG


Gas density at conditions: ρG
Gas viscosity at conditions: µG
Cut-off droplet size entrained in gas outlet: dm
Oil volumetric flow rate at conditions: Qo
Oil phase residence time required: θo
Oil surge time in oil surge compartment: θS
Oil density at conditions: ρo
Oil viscosity at conditions: µo
Water volumetric flow rate at conditions: QW
Water residence time required: θW
Water density at conditions: ρW
Water droplet in oil phase, cut-off size: dW

1. Given the cut-off size d m , calculate the terminal velocity of the liquid drop falling
through the gas phase as shown in Figure 3.17:

The cut-off size d m of the liquid droplets entrained in the gas leaving the separator
is selected by the designer to meet the design objective. This is the cut-off size
(usually in the range of 100 to 500 microns) for gravity separation. Capturing
smaller droplets than this range, if desired, can be accomplished using secondary
or enhancement devices (demister pads, coalescing vanes, etc) which intercept the
outgoing gas after it has traversed its residence volume in the separator vessel.
Sometimes coalescing vanes are placed in the path of the flowing gas phase in the
residence volume. The performance of these enhancement devices is measured by
the extent to which they reduce the effective residence time required for a given
cut-off droplet size. For example, a 50 micron cut-off is desired. By the use of
primary (gravity) separation alone a residence time of 15 seconds is required.
With an enhancement device (demister pad, for example) incorporated in the
design, the effective residence time is 3 seconds, for example. We can then say
that the enhancement device has an effectiveness of 5; it shortens the residence
time by a factor of 5. However, the effectiveness of an enhancement device is not

48
easily quantifiable; and for it we can only rely on the vendor’s data, statements or
tests.

FD : drag force ρ GVt d m


Re =
FB : net w eight of drop in gas phase µG
C D : drag coefficient 24 3
CD = + + 0.34
A : cross section of drop Re Re 0.5
(transverse to fall) V2
Re : Reynolds number of drop falling FD = C D Aρ G t
2
d m : drop diameter (cut-off-size)
πd 3
ρ L : drop density at conditions FB = (ρ L − ρ G )g m
6
ρ G : gas density at conditions
Vt : drop terminal velocity Equating FD and FB :

FD  ρ − ρG  4d m
Vt =  L  g
Vt  ρG  3C D
FB
(a) Liquid drop falling in gas phase; (b) Governing equations for
Vt is reached when FD = FB calculating Vt

Figure 3.17:
Terminal velocity of liquid droplets fallingin gas phase

At a given cut-off droplet size specified for the design, we obtain the terminal
velocity Vt of the droplet falling down in the gas phase, as outlined in Figure 3.17.

If a coalescing enhancement device (vanes, demister pads, etc) is used, we


increase Vt accordingly. This increased or effective Vt is used for sizing. For
example: Vt = 0.3 m/s according to the calculation in Figure 3.17; if the particular
device used shortens the residence time by a factor of 3, the effective terminal
velocity Vt = 0.9 m/s in calculating the gas constraint factor L*G D . See Step 3
below.

2. Given the desired oil phase residence time θ o and cut-off water droplet size
d w entrained in outlet oil, calculate ho max , the maximum allowable oil layer
thickness. (See Figure 3.18.) The cut-off water droplet size d w depends on the
dryness desired of the separated oil. For gravity separation with out any
enhancement device, the practical range of this cut-off size is 500 t0 1500 μm.
The specified residence time θ o for the oil phase is to be take as volume of the oil
phase/flow rate of the oil phase.

49
3. Select diameter D and weir height hL , determine the gas constraint factor and
liquid constraint factor as follows:

Gas constraint factor = L*G D


* 2
Liquid constraint factor = LL D

Where, L*L , L*G are, respectively, the minimum lengths of gas and liquid
longitudinal flow paths (i.e. paths parallel to the axis of the horizontal vessel) to
meet the required liquid phase and gas phase residence times.

u t2
FD = C D a w ρ o FD
2
ρ o ut d w
Re = ut
µo
For w ater drop falling in oil Re << 1 ,
FB
C D and FD can be w ritten explicitly as :
Water droplet falling in
24 oil-continuous phase
CD = → FD = 3πµo d w u t
Re
πd 3 C D : drag coefficient
FB = (ρ w − ρ o )g m
6 FD : drag force

Equating FD and FB gives : FB : net w eight of w ater drop in oil

u t : terminal velocity of w ater drop


d2
u t = (ρ w − ρ o )g w
falling in oil
18µ 0 a w : cross section of w ater drop
(transverse to fall)
The maximum allow able oil layer thickness
d w : w ater drop diameter
ho max is therefore: (cut-off size)

d θo
2 θ o : desired oil phase residence time
ho max = u tθ o = (ρ w − ρ o )g w

18µ o
Figure 3.18:
Maximum allowable oil layer thickness at desired oil phase residence time
and cut-off water droplet size

Circular Geometry

Let’s make a small digression to present the some essential elements of circular
geometry to be use in the subsequent steps. (See Figure 3.19.) We define the
ratios α and β as:
h
α= L
D

50
2
A 1  2h  2  2h  h  h 
β = L = cos −1 1 − L  − 1 − L  L −  L  ; where AC is the total
AC π  D  π D  D D
vessel cross section given by AC = πD 2 4 .

AG
ho Ao
D

hL hW AW

AL = Ao + AW
AC = AG + Ao + Aw
Figure 19:
Circular geometry: flow areas and liquid layers

Gas Constraint Factor L*G D :

Equating the time of droplet fall with the gas phase residence time:

L*G D(1 − α ) 4QG


= , VG being the gas phase velocity given by VG =
VG Vt (1 − β )πD 2
(1 − α )4QG
L*G D =
(1 − β )πVt

Liquid constraint factor L*L D 2 :

From the required water and oil residence times θ W = Aw L*L QW and θ o = Ao L*L Qo ,
we obtain:
Q θ +Q θ πD 2
AW + Ao = W W * o o , where AW + Ao = βAC = β
LL 4
4(QW θ W + Qoθ o )
L*L D 2 =
βπ

From the selected D, calculate L*L and L*G from the above expressions. The greater
of L*L and L*G will be governing in determining the length of the vessel.

4. Based on the D and hL selected, determine the oil layer thickness ho and compare it
to the maximum allowable ho max obtained in Step 2. If ho is greater than ho max ,

51
select a larger D for the vessel diameter and iterate steps (3) and (4), until ho is
smaller than ho max .

With hL selected, we have:

D2  2h   D 
AL = cos −1 1 − L  −  − hL  DhL − hL2
4  D  2 
QW θ W
AW = AL
QW θ W + Qoθ o

From AW , we obtain hW by iteration from the following implicit relation:

D2  2h   D 
AW = cos −1 1 − W  −  − hW  DhW − hW2
4  D  2 

Whence: ho = hL − hW

Elliptical Geometry

2:1 SE D (= 2r )
head
HL HL

LL LL h2
h1
LS
VS : volume betw een HL and LL
LS : surge compartment length
to meet VS
Figure 3.20:
Oil compartment surge volume VS

The volume change VS between liquid levels h1 and h2 in the surge compartment of
a horizontal cylindrical vessel with a 2:1 SE (spherical elliptical) head is given by:

π  1  π  1 
VS =  h22  r − h2  − h12  r − h1  +
4  3  4  3 

 2 −1 r − h2   r − h1 
r cos − (r − h2 ) 2rh2 − h22  − r 2 cos −1 − (r − h1 ) 2rh1 − h12  LS
 r   r 

52
The terms in the first square brackets in the above expression represent the
volume change of the liquid held in the 2:1 SE head of the vessel; the terms in the
second square brackets represent that of liquid held in the cylindrical (or
“straight”) part of the compartment. (See Figure 3.20.)

5. Based on the selected diameter D, (which satisfies the constraint of maximum oil
layer thickness in Step 4, and given the surge requirement of the oil compartment,
determine the length L S of this compartment.

The surge volume VS is the volume between LL (low level) and HL (high level) as
designated by the designer, for control and safe operation. With the elevations of
LL and HL being h1 and h2 , respectively, the surge volume is therefore VS given by
the above expression.

Given the surge time θ S and oil throughput Qo , both given as design parameters,
the length LS of the oil compartment must satisfy VS given by the relation:
VS = Qoθ S .

With L*L , L*G and LS now known, select an overall length (tangent-tangent) LTT so
that LL , LG and LS are provided as path lengths for the liquid and gas flows and as
length of oil compartment, respectively, to satisfy liquid or gas constraint,
whichever is governing, as well as the surge volume in the oil compartment.

Note that only the greater of L*L or L*G is governing; satisfying one will
automatically more than satisfies the other.

An acceptable slenderness ratio L TT /D should be obtained in consideration of any


constraint that may be posed by fabrication, transportation and plot plan;
otherwise iterate by repeating steps 3, 4 and 5.

3.6.2 Vertical Three-Phase Separator

Design Considerations

The dimension of the vessel (diameter x height) should be such that (Figure 3.21):

1. The gas phase ascent velocity must be lower than the terminal velocity of liquid
droplets at the cut-off diameter falling down through the ascending gas flow;
2. The volume of the resident oil phase is sufficient to give the required oil phase
residence time;

53
3. The volume of the resident water phase is sufficient to give the required water
phase residence time;
4. The oil phase ascent velocity must be lower than the terminal velocity of water
droplets at the cut-off diameter falling through the oil phase; and
5. An annular oil trough should be used to avoid channelling of the ascending oil
phase.

Procedure

The process parameters for the design are:

Gas volumetric flow rate at conditions: QG


Gas density at conditions: ρG
Gas viscosity at conditions: µG
Cut-off droplet size entrained in gas outlet: dm
Oil volumetric flow rate at conditions: Qo
Oil phase residence time required: θo
Oil density at conditions: ρo
Oil viscosity at conditions: µo
Water volumetric flow rate at conditions: QW
Water residence time required: θW
Water density at conditions: ρW
Water droplet in oil phase, cut-off size: dW

Gas/ Gas/
vapour vapour

PC
Gas Break D+6" PC

(42" min.) Gas Break- D+6"


-out Pan
Feed out Tray (42" min.)
Feed nozzle Feed
centreline Feed nozzle
LC centreline
LC
Oil Oil
Annular Annular
ho ho
oil trough oil trough
Interface LC
Interface
LC

hw hw
Launcher Launcher

Water
Water

(a) Gas break-out using pan; (b) Gas break-out using chimney tray;
annular oil collection trough annular oil collection trough

Figure 3.21:
Vertical three phase separator

54
1. With the cut-off size d m of droplets entrained in the gas phase, determine the
terminal velocity Vt of liquid droplets falling down through the gas phase (Figure
3.22):

Vt 2
FD = C D Aρ G FD : drag force
2 FB : net w eight of drop
πd m3
FB = (ρ L − ρ G )g C D : drag coefficient
6
ρ GVt d m A : cross section of drop
Re = (trsnsverse to fall)
µG Vt : terminal velocity
24 3 Re : Reynolds number of drop falling
CD = + + 0.34
Re Re 0.5
FD
Equating FD and FB :
Vt
 ρ − ρG  4d m FB
Vt =  L  g
 ρG  3C D Liquid droplet falling in gas phase
Figure 3.22:
Vertical Separator: liquid drop terminal velocity falling in gas phase

2. Calculate the terminal velocity u t of water droplets falling in oil phase at the
desired cut-off droplet size d W . (See Figure 3.23.)
u t2
FD = C D aW ρ o C D : drag coefficient
2
ρ o ut dW FD : drag force
Re =
µo FB : net w eight of w ater drop in oil
For w ater droplets falling in oil phase:
u t : terminal velocity of w ater drop
24
Re < 1 ; C D = falling in oil
Re aW : cross section of w ater drop
Then can be w ritten explicitly as :
(transverse to fall)
FD = 3πµo d W u t d W : diameter of w ater drop
πd 3
FB = (ρW − ρ o )g W

6
FD
Equating FD and FB gives :
Water droplet falling
d2 ut in oil-continuous phase
u t = (ρ − ρ )g W FB
18µ o
Figure 3.23:
Vertical separator: water drop terminal velocity falling through oil phase

55
3. Select a diameter D for the vessel and calculate the gas ascent velocity VG and oil
ascent velocity Vo . These ascent velocities must be lower than the respective
terminal velocities of droplets:

4
Gas ascent velocity: VG = QG ≤ Vt
πD 2

4
Oil ascent velocity: V o = Qo ≤ ut
πD 2

4. From the diameter selected, calculate the oil phase height ho and water phase
height hW , as well as overall height of the vessel. See Figure 3.21. Iterate (Steps 3
and 4) if necessary to obtain reasonable slenderness ratio (dictated by such
considerations as transportation, equipment layout and plot plan) for the vessel.

4
Oil phase height: ho = Qoθ o
πD 2
4
Water phase height: hW = QW θ W
πD 2

3.6.3 Two-Phase Horizontal Separator

Design Principle

The vessel dimension (length x diameter) should be such that (see Figure 3.24):

(1) There is enough residence time in the gas phase for the liquid droplets at the cut-
off size falling through the gas phase to reach the liquid surface;

(2) There is enough liquid volume to meet the desired liquid residence time for
residual gas in it to evolve into the gas space.

Procedure

The process parameters for the design are:

Gas volumetric flow at conditions: QG


Gas density at conditions: ρG
Gas viscosity at conditions: µG
Liquid droplet cit-off size in gas phase: dm
Liquid volumetric flow at conditions: QL

56
Liquid phase residence time required: θL
Liquid density at conditions: ρL
Liquid viscosity at conditions: µL

Feed Gas/vapour

Vt
LG
D

hL LL

Liquid
LTT
Figure 3.24:
Horizontal two phase separator

1. Given the cut-off droplet size dm, determine the terminal velocity of the liquid
droplets gas phase, as shown in Figure 3.25.

Vt 2 FD : drag force
FD = C D Aρ G
2
πd m3 FB : net w eight of drop
FB = (ρ L − ρ G )g C D : drag coefficient
6
ρ GVt d m A : cross section of droplet at the
Re = cut-off diameter
µG
Vt : terminal velocity
24 3
CD = + 0.5 + 0.34 Re : Reynolds number of drop falling
Re Re
FD
Equating F and F :
B D Vt
 ρ − ρG  4d m FB
Vt =  L  g
 ρG  3C D Liquid droplet falling in gas phase

Figure 3.25:
Horizontal separator: terminal velocity of liquid drop falling in gas phase

2. Select a vessel diameter D and liquid height hL , calculate the gas constraint factor
and liquid constraint factor as:

Gas constraint factor: L*G D


Liquid constraint factor: L*L D 2

57
We define the geometric ratios α , β as follows:

2
h A 1  2h  2  2h  h  h 
α = L ; and β = L = cos −1 1 − L  − 1 − L  L −  L 
D AC π  D  π D  D D
Where:
AC : total vessel cross section
AG
AC = AG + AL = πD 2 4
D
AG: flow area of gas phase AL
hL
AL: flow area of liquid phase

Gas constraint factor


Equating the time of droplet fall with the gas phase residence time, we have:

L*G D(1 − α ) 4QG


= ; VG being the gas phase velocity given by VG =
VG Vt (1 − β )πD 2

L*G D =
(1 − α ) 4QG
(1 − β ) πVt

Liquid constraint factor

From the liquid phase residence time θ L = AL L*L QL we obtain:

4QLθ L
L*L D 2 = , where AL = βAC
βπ

3. Determine the vessel length L TT based on L*L or L*G , whichever is governing.

Some design guide recommends the following:


4
LTT = L*G + D or LTT = LL , whichever is governing.
3

The vessel length LTT so obtained tends to be over-conservative at times. A tighter


estimate may be obtained by considering inlet and outlet nozzle locations. (See
Figure 3.26 25.) Allowing the distance S to be 3/2 times nozzle inside diameter
usually yields adequate repad-to-seam weld clearance, where S is the distance
between the gas outlet nozzle centerline and the vessel end tangent. Attachments

58
for demister pad for the outlet nozzle, if required, should also be taken into
account in determining S.

Repad-seam w eld
clearance = 2 inch approx.

Nozzle Seam
repad

S Tangent
Figure 3.26:
A more rigorous consideration for vessel length

4. Check that the vessel slenderness ratio is satisfactory for equipment layout,
fabrication and transportation. Otherwise select another D and iterate (repeat
Steps 2, 3 and 4).

3.6.4 Vertical Two-Phase separator

Design principle

The vessel dimension (length x diameter) should be such that (see Figure 3.27):

1. The gas ascent velocity in the vessel is lower than the terminal velocity of liquid
droplets at the cut-off size falling in the gas phase

2. There is enough liquid volume to meet the required liquid residence time for the
residual gas in the liquid to evolve into the gas space.

Procedure

1. Determine the terminal velocity Vt of the liquid drop at the cut-off size falling
through gas phase as shown in Figure 3.28.

2. Select a diameter D for the vessel and the operating level h of liquid; check these
selections against the gas constraint and liquid constraint such that:

59
4QG πD 2 h
VG = ≤ V t and θ L ≤ ; where VG is the gas phase ascent velocity and
πD 2 4QL
θ L is the desired liquid phase residence time; or:

4QG
By gas constraint, D 2 ≥
πVt
4QLθ L
By liquid constraint, D 2 h ≥
π

3. Dimension the vessel using guideline shown in Figure 3.27.

Gas
or vapour Vt 2 where:
FD = C D Aρ G
2 FD : drag force
FB : net w eight of drop
πd m3
D + 150**
Demister pad FB = (ρ L − ρ G )g
(150 usually) 6 CD : drag coefficient
or 170 min.

Feed A: cross section area of


Vt 24 3 drop transverse to flow
CD = + + 0.34
610 min. Re Re 0.5 Vt : terminal velocity
Re : Reynolds number of
D h Equating FD and FB : drop falling
FD
 ρ − ρG  4d m
Vt =  L  g
 ρG  3C D FB
Liquid
Figure 3.28:
**All dimensions in mm
Vertical separator: liquid drop terminal velocity falling
Figure 3.27:
through gas phase
Vertical two phase separator

C.7 Vessel geometry and antiquated design practices

In many engineering houses and technology vendors’ organizations, sizing of the


horizontal separator is based on the fixed liquid level of half a diameter; or, equivalently,
the vapour space is always allocated half the residence volume of the vessel. (See Figure
3.29.) This sizing criterion is immutable, regardless of the vapour flow rate relative to
liquid flow rate. In some applications, the relative vapour rate is low and non-governing.
The vessel is then sized to meet the requirement of liquid residence time with the fixed
half of the vessel volume. Conversely, if the liquid rates are non-governing, the vessel is

60
sized to meet the gas phase residence requirement with exactly half the vessel volume.
This practice results in an inordinately large vessel, with volume wasted on the vapour
space or liquid residence volume, as the case may be. The origin of this practice harks
back to the era of manual calculations performed with tools like manual calculators (slide
rule; hand-cranked adding machine; electric calculator), tabulated data (of circular and
elliptical geometry, etc), charts and nomographs.

w here:
AG : flow area of gas phase
AG
D AL : flow area of liquid phase (oil + w ater)
ho AO AO : flow area of oil phase
hL hW AW AW : flow area of w ater phase
D : vessel diameter
D
Fixed hL =
2
1 πD 2
Fixed AL = Ao + AW =
2 4
Figure 3.29:
Simplification by fixing liquid level hL to half D

For the horizontal vessel the exact expressions for flow area and liquid volume are given
below:

Circular part of the vessel

2
AL 1  2h  2  2hL  hL  hL 
= cos −1 1 − L  − 1 −  − 
AC π  D  π D  D D
2
AW 1  2h  2  2h  h h 
= cos −1 1 − W  − 1 − W  W −  W 
AC π  D  π D  D  D
πD 2
AC =
4 AG

AO = AL − AW AO D
AG = AC − AL hL
AW
hW

Where:
D : Vessel diameter
AC : Vessel cross section area
AG : Gas phase flow area
AO : Oil phase flow area

61
AW : Aqueous phase flow area
hL : Liquid surface height
hW : Oil-aqueous interface height

Elliptical part of the vessel

π D h 
VE = hL2  − L 
4 2 3 
L

D 2
 2h   D  
VS =  cos −1 1 − L  −  − hL  DhL − hL2  L
 4  D  2  
2:1 SE End
D
VT = VE + VS
hL
Where:
hL : Liquid surface height VS VE
D : Diameter of vessel
L : Length of the straight or cylindrical part
VE : Volume of liquid in the 2:1 spherical elliptical head
VS : Volume of liquid in the cylindrical part
Although these exact expressions would enable the designer in the past to calculate flow
areas and liquid volumes corresponding to any liquid height, the tedium involved in using
them was considered unjustifiable. The design calculations were then simplified by fixing
the liquid surface at exactly half the diameter – i.e. fixing hL at exactly D 2 . Charts,
graphs and tabulated data were developed and established on this basis; design
procedures were also established on this same basis; all these became entrenched and
immutable in the design practices of the organizations in question.

But, for more than two decades, we have had ready access to the personal computer
equipped with versatile and powerful spreadsheet and other programmable software. The
tedium that these exact expressions once, decades ago, presented to manual calculations
has long since vanished with the advent of this commonplace computing technology. One
would therefore be inclined to think that it must be a simple matter to instruct the
designer to be more rigorous to avoid wastage in equipment sizing. Besides, the use of
these exact expressions would spare him the labour of reading and interpolating charts,
nomographs and tabulated numbers – a very time-consuming, mind-numbing, repetitive
chore that is sure to cause boredom, increase the chance of mistakes and make life
unpleasant in general. No, it is not a simple matter. The designer or his organization is
wont to resist your instruction with the following arguments:

1. The first argument runs on the philosophical ground, replete as it is with


commercial and contractual implications. The engineering firm or technology
vendor has been contracted to provide the design. Tacitly, or sometimes

62
expressly, the client by awarding the contract has “bought” into the contractor’s
methods and procedures, generally claimed to be and valued as its advantageous
in-house “knowhow” and “work process”. These methods and procedures have
been in use for decades; they are parts of the time-honoured tradition of the firm
in question. An instruction to do things differently from the established methods
and procedures would constitute a change not contemplated or provided for in the
contract; and would therefore trigger changes in budget and schedule of the
project. In addition, there is always a risk to be taken in stepping out of the “tested
and proven” technology and work process.

2. The second is technical in nature. It may be true that the vapour space resulting
from always using only half the vessel volume for liquid residence time is
excessive in some situations. But the extra vapour space means that one gains
surge capacity for the liquid and this is good for operation and control. So there is
benefit after all in adhering to the tested and proven.

We know that the “technical” argument doesn’t hold water, even though it is usually
presented with a straight face; and may even be unwittingly accepted by the client. The
operating liquid surface level in the residence compartment of a three phase separator is
fixed – because it is fixed by the oil overflow weir height. The residence compartment
provides residence time and, by dint of this function, has no surge capacity, regardless of
the operating (fixed) liquid level. The separator does not and cannot function as a surge
drum, contrary to popular but mistaken belief. In other words, the separator has zero
surge capacity regardless of how generously provided the volume of the vapour space
may be. To be sure, surge volume has to be provided in the aqueous phase collection
compartment of the separator. But, here again, the fortuitous generous vapour space
actually limits the headroom for the liquid surge to go upward beyond half the diameter
of the vessel. This forces the collection compartment to compensate for this limit by
increasing the length and/or diameter of the vessel, thus compounding the wastage of
vessel volume.

But the first argument turns out to be a powerful one – the contractor’s trump card, as it
were. The mere mention of budget increase and schedule delay is enough to send a chill
down the spine of any project manager. “Technology step-out” and “new work
procedures”, unanticipated and introduced to the project on the fly, will for sure freeze
his spinal cord – you don’t run a multi-billion dollar project to test out anyone’s, not even
your own, “new” work process or new calculation methods.

One can only surmise that these arguments are compelling to many, because this
antiquated practice still abounds and by all accounts will remain immutable for a while,
wasting the client’s money.

63
Chapter 4:

OIL SKIMMING IN A HORIZONTAL SEPARATOR

4.1 Introduction

In many chemical and hydrocarbon processes aqueous streams are produced that have in
them dissolved gas and entrained oil. Sour water and rich amine streams are notable
examples. The dissolved gas and entrained oil have to be removed before these streams
can be further processed. This is usually accomplished in a horizontal three-phase
separator fitted with the appropriate internals for oil skimming. Oil skimming, as opposed
to bulk oil removal, is used because the oil present in this application is incidental and in
indeterminate but small quantity.

For the presentation in this chapter, we shall work with a rich amine flash drum, which is
the quintessential application of a horizontal separator with oil skimming. In fact we shall
use the terms “horizontal separator with oil skimming”, “amine flash drum” and
“skimmer” interchangeably to mean the same thing. We shall discuss the common and
almost universal design flaws in most commercial offerings of the amine flash drum and
their consequences in operation.

4.2 Amine Flash Drum

The rich amine flash drum is used in petroleum refineries and other hydrocarbon
processing installations, such as heavy oil upgrading and gas processing plants. In a
typical installation, lean amine, which is an aqueous solution of an amine and is relatively
free of any absorbed species, is circulated to various units or plants where it is used to
absorb acid gas (H 2 S and CO 2 ) by liquid-vapour contacting in a tower with trays or
packings known as the contactor or the absorber. The amine so loaded with absorbed acid
gas is called rich amine. Apart from the absorbed acid gas, which is weakly chemically
bonded to the amine molecules, the aqueous amine solution is also saturated with
dissolved light hydrocarbon gases and may also have heavier liquid hydrocarbons (oil)
entrained in it. The presence of entrained oil is incidental, which means in theory it
shouldn’t be present in the rich amine. In practice, however, it is there – in small and
indeterminate quantity. The entrained oil must be removed from the rich amine before it
is regenerated by heat and fractionation to break the weak acid gas-amine bond, to yield
lean amine, which is then circulated back to the users, to complete the loop. (The acid gas
captured in this regeneration process is routed to other units for further processing.)

In most amine regeneration facilities, charcoal (activated carbon) is used to remove by


absorption organic materials including oil from a slip the amine solution, in a piece of
equipment known as “charcoal filter”. But the charcoal filter is only a polishing measure,
removing the last traces of oil; most of the entrained oil must be removed before the rich
amine reaches the regeneration facility. The presence of oil in amine causes foaming in
the regenerator and the contactors, and rapid exhaustion of the polishing filters. The

64
malfunctioning of the regenerator in turn causes major process upset both within the
amine regeneration facility and in plants that receive the regenerated lean amine.

Rich amine streams are collected from various users. Rich amine streams from high
pressure users are usually flashed down to some intermediate pressure to release most of
the dissolved gas before they are collected in the rich amine return header. The collected
rich amine is delivered to the amine flash drum, which operates at near atmospheric
pressure. The remnant gas in the rich amine flashes off in this flash drum while the
quiescent gas-free liquid is rich amine solution with an incidental small amount of oil
entrained in it. The rich amine flash drum is a horizontal three phase separator, which
provides residence time for the entrained oil to separate by gravity and which has
internals arranged to skim off and collect the separated oil.

4.3 Design of Amine Flash Drum

LTT
Inlet LG
Gas /
LW vapour
hA : aqueous overflow w eir height

y , oil layer thickness


LS hL : oil overflow w eir height
1
: oil-w ater interface level
2
H D
hW
hW hL : approach velocity in crest
hA Vo
Underflow (aqueous overflow horizontal velocity)

H : crest height
(a) Major vessel dimensions
Oil Aqueous
(Amine) B : aqueous overflow w eir w idth
Oil w eir (trough edge)A
Vo2 : aqueous overflow kinetic head
Oil trough 2g
Vo
H
H

hL Underflow
w eir
Upflow

B
hU
hA
Underflow hA
WUP Overflow w eir
View A - A
A

(b) Underflow weir and upflow channel width (c) Overflow weir height and width

Figure 4.1:
Skimmer: Dimensions and nomenclature

It is noteworthy that, the amine flash drum, a horizontal three-phase separator with oil
skimming and collection, is quite different in its operation from the horizontal three-
phase separator for bulk oil-water separation. (See Section 3.3.) In the amine flash drum

65
the oil-water interface is not controlled. Instead, an expected oil layer of a certain
(calculated) thickness is maintained. The oil layer overflows into a collection trough
called the oil trough from which is it pumped out, usually intermittently. In the case of
bulk oil-water separation, the oil-water interface is controlled by adjusting the aqueous
phase draw-off rate. (See Figure 3.4.)

The principle governing the skimmer design is the energy conservation (Bernoulli)
equation applied to the flow path between point 1 and point 2 (Figure 4.1):

Vo2
ρ A (hL − y )g + ρ o yg = ρ A (h A + H )g + ρ A + ρ A hF g (4.1)
2

Where: ρ A , ρ o are the mass densities of the aqueous and oil phases, respectively; and hF
is the head loss due to friction and abrupt change in flow direction through the underflow
and upward to the overflow. (See Figure 4.1 for the meaning of the other terms in the
equation.) Noteworthy is the crest of the overflowing liquid formed on top of the
overflow weir. This crest has a significant height H, significant in the context of the
design equation (4.1). Appendix IV shows how this design equation for the skimmer is
derived from the first law of thermodynamics, the law of the conservation of energy.

In most commercial designs offered by the major engineering houses and technology
vendors the existence of the crest height is ignored. The existence of crest height is also
denied by at least one published design manual (see Arnold & Stewart (4)) and in all the
in-house design procedures of the engineering houses and technology vendors that the
present author has come across. Also ignored by these designers are the kinetic head
Vo2 2 g of the overflowing liquid as well as the head loss term ρ A hF g . The kinetic head
Vo2 2 g is of the order 20 to 50 mm; and this is not negligible in the context of the present
application. See Appendix V for the design procedure of the three phase separator with
oil skimming.

4.4 Consequences of the Wrong Design

As basis for the discussion in the rest of this chapter, we use the process parameters taken
from an actual design of an amine flash drum; these parameters are:

Amine flowrate Q 0.161 m3/s (580 m3/h)


Amine overflow weir height hA 3120 mm
Oil overflow weir height hL 3236 mm
Drum diameter D 4000 mm
Underflow weir height hU 500 mm
Upflow channel width WUP 500 mm
Amine density ρA 1028 kg/m3

66
Oil density (weathered) ρo 850 kg/m3

This is a sizable amine flash drum with a throughput of 580 m3/h (2553 GPM) of rich
amine collected from various processing units in a large heavy oil upgrading complex.

In many situations we deliberately use simplified equations and approximate relations in


our calculations. We do so with the full knowledge that what is knowingly ignored in the
process of simplification or approximation has no significant bearing on the end results,
which results may be, for example, equipment size, motive power or internal details of a
vessel. The very energy conservation (Bernoulli) equation (4.1) we use here is itself such
a deliberate simplification. What we knowingly ignore, for instance, in the Bernoulli
equation used here are the kinetic head of the liquid and head loss incurred by the liquid
flow in the residence compartment of the separator, as well as the kinetic head of the oil
flowing into the oil trough. To ignore judiciously and be fully aware of the resulting
insignificance is one thing; it is, however, quite another to ignore significant terms out of
ignorance, and this can and often does lead to unpleasant surprises!

So, the following unjustifiably truncated equation is used almost universally for the
design of the horizontal three phase separator with oil skimming:

ρ A (hL − y )g + ρ o yg = ρ A (h A + 0)g + 0 + 0 (4.2)

The quantities ignored are marked as 0, for comparison with the appropriate design
equation (4.1). It can be seen from this truncated equation that weir height difference
(hL − h A ) calculated from it will be far too low. For example, based on the desired
operating oil layer thickness of y = 100 mm, the elevation difference (hL − h A ) between
the oil overflow weir (upper edge of the oil trough: see Figure 4.1) and the amine
overflow weir is, according to this unjustifiably truncated equation (4.2):

(hL − h A ) = ρ A − ρ L y , giving (4.3)


ρA
(hL − h A ) = 17 mm.
Note that in the preceding equation (4.3), none of the terms on the right-hand side
depends on the flow rate of the liquid, y being the oil layer thickness desired as specified
by the designer. To an alert and discerning designer, this independence of (hL − h A ) from
the flow rate should ring an alarm bell on the validity of this unjustifiably truncated
equation! Figure 4.2 shows the dimensions of the wrong design based on this
unjustifiably truncated equation.

Because the crest height H and the velocity head Vo2 2 g of the amine overflow are
ignored in the wrong design shown in Figure 4.2, the oil overflow weir hL height of 3137
mm turns out to be too low with respect to the amine overflow weir height h A of 3120
mm (selected by the designer). Suppose this design mistake goes unnoticed – as more

67
often than not it does - and the separator is then constructed based on the dimensions
given in Figure 4.2. In operation with this wrong design, initially oil has not accumulated
enough to form an appreciable layer. The liquid head in the residence compartment will
assume a value hL* , which is always greater than the hL intended by the designer. Based
on the process parameters given above, H , Vo2 2 g and hF , all flow-dependent, turn out to
be 73.44 mm, 22.86 mm and 1.65 mm, respectively, which are not taken into account in
the wrong design. This state of affairs is shown in Figure 4.3. In this figure, the oil trough
is flooded (with amine) and the oil pump is assumed to be running at 2.27 m3/h (10
GPM).

Initially, the liquid head in the residence compartment being all amine, the liquid level hL*
in this compartment is higher than h A by the amount:

V 2 
(h )
− h A = H +  o + hF  = 98 mm
*
L
 2g 
This is far greater than (hL − h A ) = 17 mm. This means the oil through is completely
flooded with amine solution and it will remain so as long as the vessel is in operation
with the same feed rate of rich amine. With the passage of time, as more and more oil
separates and accumulates, an oil layer forms and grows in thickness. As it does, hL*
(now the top of the oil layer) moves upward while the oil-amine interface moves
downward, as governed by the Bernoulli equation. In other words, the resident oil layer
grows with time in both directions, upward and downward as depicted in Figure 4.4.

68
As can be seen from Figures 4.5, 4.6 and 4.7, how the wrong design (with hL − h A = 17
mm) actually fails in operation depends on the height of the partition baffle the designer
happens to set. The partition baffle height is immaterial to the normal operation of a

69
correctly designed skimmer provided it is set somewhat higher than the oil weir h L . It
may be set higher than h L by 50 to 150 mm, a margin deemed adequate to prevent oil
from overflowing (or jumping over) the partition baffle due to wave action in the
residence compartment. Thus the partition baffle height is set at the discretion or
preference of the designer or according to some guideline used in his design procedure.
In what follows we look at three cases in which the wrong design fails, depending on the
height of the partition baffle the designer happens to have set.

Case A: Partition baffle height set at 3230 mm

In Figure 4.5, the partition baffle height is set at 3230 mm by the designer. Here the oil
surface hL* has risen enough to overflow the top edge of the partition baffle into the amine
surge compartment. A stable operating oil layer is maintained. The oil trough is
completely flooded with amine solution; no oil can collect into the oil trough. In
calculating the dynamics shown in Figure 4.5, the oil disposal pump is assumed to have
been switched on by high level sensed in the oil trough, pumping out at the continuous
rate of 2.27 m3/h (10 GPM). As the pump-out rate from the oil trough is small (compared
to the amine throughput), it can be seen that, if the oil trough pump is not running, the
failure shown in Figure 4.5 still prevails, with the oil layer slightly thinner than 70 mm.
Here we have an absurd situation, where the oil is separated in the residence
compartment only to be blended back into otherwise oil-free amine in the amine surge
compartment!

70
Case B: Partition baffle height set at 3185 mm

In Figure 4.6, the partition baffle height happens to be set at 3185 mm. Here amine
solution overflows the partition baffle continuously. Unlike Case A (Figure 4.5),
however, at this partition baffle height an oil layer cannot be formed in the residence
compartment; any separated oil is swept away (as a sheen) on the surface of the amine
solution overflowing the partition baffle. The outcome is the same as Case A: all
entrained oil in the rich amine feed ends up in amine surge compartment and leaves the
separator with what is supposed to be an oil-free rich amine stream.

Note that Figure 4.6 shows the oil trough pump as running. Because of the small pump-
out rate, the failure is practically the same with or without the pump running, with only
insignificant differences in the crest heights of the overflowing amine. Also, the failure
depicted in Figure 4.6 prevails for any height of the partition baffle set lower than 3185
mm, with the amine overflowing the partition baffle at a rate greater than 98.27 m3/h and
a correspondingly lower rate of amine underflowing into the surge compartment.

Vessel w all Vessel ID = 4000


Oil trough Partition baffle set at 3185
completely flooded 3
98.27 m /h Amine overflow ing
H = 64.72 partion baffle w ith
y = 0 (oil layer can't form) Vo2 2 g= 20.19 crest = 22.21
Overflow velocity head = 7.48
hL = 3137 3
479.46 m /h
Ah = 3120
*
h L= 3207 Amine level
Upflow

3 Oil trough
579.89 m /h Amine Surge
level
Compartment LS
Residence Underflow
Compartment Underflow w eir 2.27 m3/h
(10 GPM)

Underflow height hU = 500


Upflow w idth WUP = 500
Under/upflow friction hF = 1.65

Figure 4.6: Case B : Partition baffle height set at 3185 mm


No sizeable oil layer can form; amine overflowing the partition baffle at
98.27 m3/h with sheen of oil on it.

Case C: Partition baffle set 3240 mm

This case is shown in Figure 4.7. A layer of oil is held in the residence compartment; but
the partition baffle is high enough to prevent the oil from overflowing its edge into the
amine surge compartment. The oil trough is permanently flooded and the pump-out is

71
continuous at the oil disposal pump capacity of 2.27 m3/h. This pump capacity is, by
stipulation, greater than the rate of oil entrained in the rich amine feed to the separator. In
normal operation, with a correct design, the oil disposal pump is intended to be switched
on and off intermittently to remove oil that collects in the oil trough. Since the continuous
pump-out rate is greater that the oil accumulation rate in the oil layer, the oil-amine
interface will hover about the edge of the oil weir. The upshot is that a mixture of oil and
amine is continuously pumped out of the oil trough.

Vessel w all Vessel ID = 4000


Top edge of partition
baffle = 3240

y = 90.31 Vo2 2 g = 22.86 Crest


H = 73.44
3
hL* = 3234 577.62 m /h
h A = 3120
fluctuating Amine
slightly Amine level
+ oil
Upflow

hL = 3137 WUP Oil trough


3 Amine Surge
579.89 m /h level
Residence Compartment LS
Compartment Underflow hU
2.27 m3/h
(10 GPM) oil+w ater
Underflow height hU = 500
Upflow w idth WUP = 500
Under/upflow friction hF = 1.65
Oil-amine interface fluctuates to let amine slip into oil trough

Figure 4.7 Case C: Partition baffle set at 3240 mm


With continuous oil trough pump-out at 2.27 m3/h, oil-amine interface
hovers at the oil weir; as pump-out rate is greater than oil
accummulation rate, the oil-amine interface fluctuates to let amine slip
into the oil trough and get pumped out with oil.

4.5 A Correct Design

Using the same process parameters as before and a desired operating oil layer of y = 100
mm, the appropriate design equation (4.1) gives us the correct design with hL − h A = 116
mm; not 17 mm according to the unjustifiably truncated equation (4.2). An operating oil
layer thickness of y = 100 mm is considered adequate to withstand perturbations without
the oil layer getting obliterated; such perturbations may be caused by wave action in the
residence compartment and changes in crest height and amine overflow kinetic head
brought about by throughput fluctuation. For example, the oil layer thickness decreases
with an increase in amine throughput; with design oil layer thickness of y = 100 mm, it is
expected that an increase in amine throughput will not wipe out the oil layer and causes
amine to over flow into the oil trough. Figure 4.8 shows the dynamics of the skimmer
operation with these process parameters.

72
Note that the operating crest height H =73.63 mm and the amine overflow velocity head
Vo2 2 g = 22.92 mm are properly accounted for in the energy balance by the use of (4.1).
Note also that, with an underflow weir height hU = 500 mm and upflow channel width
WUP = 500 mm, the head loss is kept small at hF =1.65 mm. Usually, when these two
dimensions, hU and WUP , are set at 12 to 15 % of the vessel diameter, the friction loss hF
should turn out to be below 3.00 mm.

Vessel w all Vessel ID = 4000

Oil in trough removed


on trough level control
Overflow
Vo2 2 g = 22.92 H = 73.63

y = 100 h A = 3120
Oil-amine WUP Oil trough
Upflow

Amine level
interface
hL = 3236 LS
level

Underflow
hU
Skimmed oil
Underflow height hU = 500
Upflow w idth WUP = 500
Under/upflow friction hF = 1.65

Figure 4.8:
Correct design with proper hL − h A = 116 mm (not 17 mm!).
Operation as depicted is real: sustained oil layer and oil
overflowing into oil trough and pumped out intermittently.

4.6 Example of a Commercial Offering

Figure 4.9 shows a partial transcript from an equipment data sheet prepared for an amine
flash drum with oil skimming. The data sheet was issued for detailed mechanical design
and subsequent fabrication. The design throughput of this amine flash drum is 1624 m3/h
(7150 GPM) of rich amine; this is a sizable vessel, as can be seen from its dimensions. It
was intended to be the main amine flash drum of a major heavy oil upgrading complex.
Note the weir height difference hL − h A = 22 mm. This wrong weir height difference is
obtained using (4.2), which ignores the crest height H , the amine overflow kinetic head
Vo2 2 g and the albeit small head loss hF incurred in the underflow and upflow channel.
The parameters used in (4.2) are: ρ A = 1028 kg/m3, ρ o = 800 kg/m3 and the desired oil
thickness of y = 100 mm. Note also that the oil weir and the partition baffle are set at the

73
same height. With the amine overflow weir height h A = 5478 mm set by the designer, the
design (intended) throughput would correspond to a crest height of H = 84 mm. As can
be easily seen that, if built according to the dimensions shown in Figure 4.9 (with h L - h A
= 22 mm), this vessel simply would never work. In operation, the amine solution would
just overflow the partition baffle into the amine surge compartment; it would underflow
into this compartment as well; the oil trough would be completely flooded with aqueous
amine; no oil separation and collection would be possible. All oil entrained in the rich
amine feed to the amine flash drum leaves the vessel with what is supposed to be an oil-
free amine stream.

Gas

Partition baffle
Feed
Oil w eir Crest height
22 (fictitious)
22
H = 84

7010 ID hL =
hA =
5500
5478

Amine
Oil Amine overflow
28042 T-T w eir

(a) Basic dimensions (b) Fictitious crest height

Figure 4.9 :
A design gone mad: hL − h A = 22 mm; the design throughput of 1624 m3/h of rich amine
feed corresponds to a crest height of 84 mm, but this crest height cannot be sustained
due to the wrong hL − h A .

4.7 Sensitivity of the oil layer thickness to the flow dynamics

It is instructive to see to re-arrange (4.1) as:

ρA   Vo2 
y= (hL − h A ) − H −  + hF  ,
ρ A − ρ o   2g 

so that we can examine the sensitivity of the oil layer thickness y to the change in the
dynamic quantities H, Vo2 2 g and hF .

Taking ρ A , ρ o and (hL − h A ) as parameters, we can take the differential of the preceding
re-arranged Bernoulli equation:

74
∂y ∂y ∂y
) (
d Vo2 2 g + )
dy = dH + dhF (4.4)
∂H 2
(
∂ Vo 2 g ∂hF

The preceding equation (4.4) expresses the variation of y due to changes in H,


Vo2 2 g and hF .

Taking the finite difference approximation of (4.4), we have:

ρA  Vo2 
∆y =  − ∆H − ∆ − ∆hF  ; noting that: (4.5)
ρa − ρo  2g 

∂y ∂y ∂y − ρA
= = =
(
∂H ∂ Vo 2 g
2
)
∂hF ρ A − ρ o

For want of a better word, let’s call the density ratio ρ A (ρ A − ρ o ) the amplification
factor. It amplifies the changes of the dynamic quantities shown in the square brackets on
the right of (4.5). At ρ A = 1.00, ρ o = 0.80, the amplification factor is 5. This means a
change, for example, in H given by ∆H = +4 mm (due to increase in throughput) causes
the oil layer thickness to change by ∆y = - 4 mm x 5 = - 20 mm. At ρ A = 1.00, ρ o = 0.90,
the amplification factor is 10! The call for rigour in the design calculations and in the
setting of design parameters cannot be over-emphasized.

Given the sensitivity of the operating oil layer thickness y to these dynamic quantities, it
is therefore preferable to keep H, Vo2 2 g and hF as small as possible by design. For
example, if the crest height by design is too large, then an increase in it (due to a surge in
feed rate) could wipe out the operating oil layer and cause amine to flow into the oil
trough. The friction loss hF can be kept small by a reasonable choice of the dimensions
for the underflow weir and the upflow channel, as discussed above. The crest height can
be kept small by the use of a large diameter for the vessel. However, other requirements
and constraints also bear on the choice of vessel diameter, such as length/diameter ratio,
transportation and fabrication, etc. Given these opposing considerations, the following
design targets are recommended:

Operating oil layer thickness, y: 100 mm


Crest height, H: 50 to 75 mm
Underflow + upflow losses, h F : < 3.00 mm

The aqueous phase (amine) overflow velocity head Vo2 2 g usually falls out to be about
20 to 30 mm within the range of the recommended design crest height. An operating oil
layer thickness of y = 100 mm is considered adequate to withstand perturbations without
the oil layer getting obliterated; such perturbations may be caused by wave action in the
residence compartment and changes in crest height and amine overflow kinetic head
brought about by throughput fluctuation.

75
For detailed requirements and design considerations, see Appendix IV.

4.8 “Performance” in the field

Like many ill-conceived designs, the amine flash drum is notorious among operators as
the “expletive thing” that never seems to work; and in reference to the liquid pumped out
from the oil trough, their comments are: “sometimes it gives us plenty of oil; sometimes
nothing but pure amine!”. Through the numerical illustrations presented above we can
see the veracity in the laments of these operators. They also say the same of the sour
water flash drum, which is identical in design principle to the amine flash drum, and of
most vessels and tanks from which they are supposed to skim off some incidental oil.

In most installations, the liquid from the oil trough is collected as “wet slop” and sent to
the wet slop tank. The wet slop tank provides residence time for the water in the wet oil
to separate and settle out. This water is regularly drained off and sent to the oily water
collection and processing system. As the wet slop tank also collects wet oil streams from
other sources, the loss of amine to the wet slop usually goes unnoticed, especially when
the other wet slop streams are overwhelmingly larger. In addition, there are usually many
other oily water streams, besides the water drained from the wet slop tank, that flow into
the same oily water collection and processing system. The amine present in the total oily
water body is just the proverbial drop in the bucket – hardly detectable. The only sure
indication of amine loss is the depletion of its inventory. The average amine
“consumption”, i.e. rate of replenishing amine inventory, in the hydrocarbon processing
industry is about three to five inventory volumes per year! Most of this loss, we surmise,
is due to inoperable amine flash drums thanks to the universal unflinching denial of the
crest on the amine overflow weir. If only someone could see this horrendous amine loss
in the proper perspective: the price of amine, volume for volume, is higher than the price
of beer!

4.9 One lone voice in the wilderness

It is interesting to note how a simple principle such as that governing the skimmer
operation can be so misunderstood or ignored. The lack of attention on the part of the
designer or the wrong design methods adopted by his organization may be to blame. The
present author has not found one engineering house or technology vendor that has the
right design equation for the skimmer, i.e. the energy conservation (Bernoulli) equation
(4.1), if it has one at all. Published process equipment design methods offer the same
unjustifiably truncated equation (4.2) for the skimmer design referred to earlier. They all
categorically deny the existence of the overflow crest H and its concomitant kinetic head
Vo2 2 g .

However, one lone voice has been heard in the wilderness. In the design manual of one
major oil company, the design equation of the oil skimmer does include the crest height,

76
but not the amine overflow kinetic head Vo2 2 g and head loss hF . These two terms were
left out knowingly; but the effects of these two terms are compensated for in the design
procedure. The said manual provides sizing guidelines for the dimensions of underflow
weir and the upflow channel to ensure a low head loss ( hF < 3.00 mm), so that it is
insignificant and can be ignored in the design equation. Also, a design oil layer thickness
of = 225 mm (instead of 100 mm) is used in the calculation procedure in the said
manual. This apparently excessive oil layer thickness by design is specified in
anticipation of the effect of the amine overflow kinetic head Vo2 2 g . When this kinetic
head comes to bear in operation, the operating oil layer will thin out in response and
should find itself operating in the range of 50 to 100 mm (not 225 mm provided for in the
design calculation). See discussion in Section 4.6 on the sensitivity of oil layer thickness
to flow dynamics. The omission of the kinetic head Vo2 2 g and the head loss h F was
made, judiciously, in consideration of the rather involved and unwieldy calculations that
that these two terms would entail in the context of the calculation tools available at the
time, before the advent of the spreadsheet and other computational tools.

77
Chapter 5:

SHELL AND TUBE HEAT EXCHANGERS

5.1 Introduction

In this chapter, we address some of the common misconceptions and disconnects between
the process designer and other disciplines engaged in the design and the fabrication of
shell and tube heat exchangers.

The design of a shell and tube exchanger involves two principal disciplines: process and
mechanical. Other disciplines are also involved: metallurgy, welding, piping and civil and
structure, and more. While all these disciplines jointly are responsible for the ultimate
integrity and functionality of the exchanger in operation, only the design considerations
and practices of the process and mechanical disciplines are discussed here as they pertain
to the geometry and construction the heat exchanger. The overlap or separation of the
design activities and of the persons involved in these two principal disciplines varies
from one organization to another. In this discussion, the generic term designer is used to
describe the person or persons involved in any of these design activities. The terms
process designer and mechanical designer may be used when the design considerations
or activities are specific to either of these two disciplines; and they are not intended to
discriminate one discipline against the other, or to differentiate and inculpate persons
engaged in the design effort. Nor is it intended to advocate that design work should
consist of fragmented undertakings of isolated disciplines.

The following points will be highlighted in the course of our discussion:

• The process specification and data input to the design of shell and tube
exchangers;

• The aspects of exchanger geometry and construction crucial to the process


functions of the exchanger, especially those process functions that entail phase
change, i.e. boiling and condensing, partial or total;

• The common pitfalls in the communication "paper work" between the process
designer, the mechanical designer and the fabricator of the exchanger; and

• The common pitfalls in the over-reliance on the part of the designer on warning
messages from heat exchanger design software.

We shall illustrate these points by going through some examples of heat exchanger
design efforts. Also, we shall discuss some common misunderstanding and design
practices regarding the following:

• Temperature pinch; and


• Abuse of cooling water temperature.

78
5.2 Baffle Cut Orientation

The importance of baffle cut orientation will be illustrated in the subsequent sections as
the various applications of the shell and tube exchanger are discussed. There are two
baffle cut orientations, vertical and horizontal, as shown in the sketches below.

One baffle space One baffle space

Bundle
axis
Bundle
axis
o
(Rotate 90 to obtain o
horizontal baffle cut) (Rotate 90 to obtain
Single-segmental horizontal baffle cut)
Double-segmental
Vertical cut
Vertical cut

Figure 5-1: Figure 5-2:


Single segmental baffles - with vertical cut. Double segmental baffles - with vertical cut.
The horizontal cut orientation can be The horizontal cut orientation can be
obtained by rotating the the tube bundle by obtained by rotating the the tube bundle by
90 degrees about the bundle axis. 90 degrees about the bundle axis.

As shown in Figures 5-1 and 5-2, the baffle cut orientation, vertical or horizontal, refers
to the orientation of the straight edge or edges of a baffle, which is a truncated circular
plate. If these edges are vertical, we say the baffle cut is vertical; if these edges are
horizontal, then the baffle cut is said to be horizontal. Note that the baffle cut orientation,
vertical or horizontal, is relevant only if the exchanger shell is installed horizontally.
Shown conceptually is the shell side fluid meandering from one baffle space to the next.

Longitudinal
Longitudinal
baffle
baffle

Single-segmental Double-segmental
vertical cut in vertical cut in
TEMA F, G and H shell TEMA F, G and H shell
Figure 5.3: Figure 5.4:
Single segmental baffles with vertical cut . The Double segmental baffles with vertical cut . The
interlacing longitudinal baffle creates two passes interlacing longitudinal baffle creates two passes
in the shell as used in TEMA shell types F, G in the shell as used in TEMA shell types F, G
and H and H

79
In Figures 5.3 and 5.4 are shown the transverse baffles that are interlaced with a
longitudinal baffle to create, in a single physical shell, two flow passes in the shell side.
This technique to obtain two flow passes in the shell side is exploited the TEMA shell
types F, G and H. (Consult any shell and tube exchanger design handbook or texts on the
designations of TEMA shell types.)

5.3 Process Input in the Data Sheet – Baffle Cut

The process requirement of heat exchanger results from the process design calculations.
These calculations yield the process data such as: temperature and pressure, both
operating and design; the states and properties of the fluids flowing through the
exchanger, flow rates and duty. The process designer transmits this process requirement
to the other disciples in the form of a data sheet for the shell and tube exchanger. The
data sheet varies significantly from one organization to another.

At this point, depending on the particular organization of his firm, the process designer
should have done enough work to indicate the following requirements of geometry and
configuration as a go-forward for the designers in other disciplines:

• The exchanger size;

• The TEMA exchanger type (AEU, for example) for the service;

• Exchanger configuration - number of shells in series and in parallel;

• Number of tube passes and shell passes per shell;

• Baffle type, orientation and spacing; and

• Nozzle arrangement and orientation.

We will look at some examples of the data sheet entries made by the process designer to
express his requirement and intention for the design of the exchanger. These examples
are transcribed from actual data sheets to show the pertinent points under discussion. The
data sheets from which these examples are extracted were actually issued for detailed
mechanical engineering and fabrication.

Example 1:

Baffles
Type: Double segmental
Cut: 26.5 % Vert. [Note: 26.5 % of diameter or area?]
Number of baffles: 11
Baffle-longitudinal: None

80
In Example 1 there is no dedicated box (or entry) in the data sheet to state the baffle cut
orientation, but this orientation (indicated by “Vert”) was “forcibly” added in at the "%
cut" entry. In other words, the person who stated vertical as the baffle cut orientation he
wanted did so forcibly, because there is no entry in the data sheet that explicitly requires
or reminds him, or anyone else, to do so. Note also that 26.5 % cut means either the
circular plate is to be truncated by 26.5 % of its area or 26.5 % of its diameter. Both
conventions are in use; and quite often simultaneously in the same organization! There is
no indication anywhere in the data sheet in question as to which convention was followed
in this particular data sheet.

Example 2:

Baffles
Cross baffle type: Double seg.
Cut: 29 % [Note: 29 % of diameter or area?]
Cut orientation: X Parallel Perpendicular [Note: Parallel to what?]
Number of baffles: 11

In Example 2 specific "check" boxes are provided to permit two choices of baffle cut
orientation: Parallel or Perpendicular; but parallel or perpendicular with respect to
what? Actually, parallel or perpendicular with respect to the centerline of the shell side
inlet nozzle, according to convention, but not many people are cognizant of this
convention! (This practice of specifying the baffle cut orientation as parallel or
perpendicular with respect to the shell side inlet nozzle centerline is normally reserved
for shells installed vertically, for obvious reason, because in a vertical shell , the
descriptive “vertical or horizontal, in reference to baffle cut orientation has no meaning.)
And nowhere in the data sheet was the orientation of the shell side inlet nozzle given; and
neither was there any reference in the data sheet as to where this piece of information
could be found. To compound the ambiguity, the subject exchanger, as it turned out, was
intended to be a horizontal one! Note also the ambiguity in % cut (of area or diameter?)

Example 3:

Baffles
Baffles-cross type: Segmental
% cut (Diam/Area): 29 [Note: 29 % of diameter or area?]
Baffle spacing: 18-5/8 inches
Baffle-longitudinal: None

In Example 3, there is no dedicated entry to state baffle cut orientation; and such
orientation was not forcibly stated or noted anywhere in the data sheet either. Note also
that the entry provides the choice to state explicitly the % cut in either diameter or area;
but in the example, even this choice was not made by the designer! Neither do we know
if the designer intended the baffles to be single or double-segmented.

81
As mentioned earlier, the above examples are extracts from actual datasheet approved for
detailed mechanical design and fabrication. Ambiguity, omissions and incompleteness
inherent in the datasheets often have unpleasant consequences. The detailed mechanical
designer, faced with this ambiguity, omissions and incompleteness, is wont to assume
that the mechanical details in questions are not relevant to the process function of the
exchanger; he would most likely take the liberty to exercise his discretion and preference.
His preference, for example, may be to have the baffle cut horizontal unless indicated
otherwise; or, again for example, all baffle cuts are single-segmented, unless indicated
otherwise.

5.4 How Geometry Can Be Crucial to Process Functioning

We will discuss three categories of applications of the shell and tube exchanger to
underscore the importance of exchanger geometry to its process functions. In the first
category, the process on the shell side entails condensation, partial or total, of the process
fluid. In the second, we highlight the geometry crucial to the workings of a reboiler with
the process fluid boiling in the shell side. In the third, we discuss the favorable geometry
of a tube side condenser. The discussion in this section is a brief outline of the crucial
exchanger geometry. More supplementary details will emerge in the subsequent Section
5.5, where we work with exchangers of specific TEMA shell type and tube bundle
construction.

5.4.1 Condensation in the Shell Side of a Horizontal Exchanger

For this application it is obvious that the baffle cut orientation must be vertical to allow a
relatively free drainage of the liquid dropped out of the condensing fluid. (See Section
5.5.4). In a shell with horizontal baffle cuts, the shell side would be flooded with liquid,
which would severely impede the condensation process. The condensing vapour cannot
access flooded tube surface to have its heat removed and to condense. Major process
upset and erratic pressure surge are the immediate signs of this malfunctioning.

There are of course weep holes at the low point of the baffles with horizontal baffle cut.
These weep holes are intended to provide complete drainage after a shutdown and steam-
out in preparation for opening the exchanger and for tube bundle removal. They are far
from adequate (and are not intended) for draining the condensate at the rate produced
during normal operation.

5.4.2 Boiling in the Shell Side of a Horizontal Exchanger

To effect boiling in a horizontal shell, the internal construction must be such that the
liquid is introduced at the bottom of the shell, where it is allowed to spread longitudinally
and relatively unobstructed before rising to contact the tube bundle. Once the spread-out

82
liquid reaches the horizontal tube bundle it rises due to the buoyancy created by boiling
as the liquid comes into contact with the tube surface.

The two phase fluid (liquid + vapour) comes to the top of the horizontal tube bundle,
where the fluid must be provided with space to flow longitudinally, with minimum
obstruction, to reach the outlet nozzle or nozzles.

This flow pattern that prevails when boiling a liquid in a horizontal shell is known as the
cross flow pattern and is contemplated in a TEMA X shell, as well as in commercial
software for the design of shell and tube exchangers. The bundle geometry must be one
that facilitates this flow pattern. (See 5.5.3 NTIW Tube Bundle.) Any attempt,
inadvertent or otherwise, to “frustrate” or obstruct this naturally established flow pattern
will result in a severely under-performing or an inoperable exchanger.

Boiling in the shell also requires that tube pitch and pitch angle are such that they
facilitate the dislodging of vapour from the tube surface: large pitch (1.75 to 2.00 times
tube OD) and a 30-degree angle of encounter.

5.4.3 Tube Side Condensation in a Horizontal Exchanger

In most applications, such as the overhead condenser of a fractionation tower, the vapour
enters the condenser saturated. This means condensation begins immediately upon
contact of the vapour with the tube surface – no desuperheating of the vapour is required
beforehand. This also means that the tube side heat transfer coefficient, condensing only,
is practically independent of vapour velocity, (unlike the single vapour phase heat
transfer without condensing where the heat transfer coefficient is strongly dependent on
vapour velocity).

This independence of velocity enables us to arrange all condensing in a single tube pass,
since we no longer need to consider a multi-pass tube bundle to achieve an acceptable
vapour velocity. The single pass bundle can be tilted to enhance the drainage of the liquid
and to facilitate contact of the condensing vapour with the tube surface.

A point often missed by the designer is that, in condensing service, efficient drainage of
the liquid formed (whether it is shell side or tube side condensing) is of great importance.
Tube surface covered by liquid is not accessible to direct vapour contact and this severely
reduces the condensing capacity. Quite often a multi-pass tube bundle oriented
horizontally is used for tube side condensing services. The loss of condensing capacity is
incurred due to flooding inside the tubes. In a horizontal tube, flooding has to occur
because the liquid has to build up a hydraulic gradient in the tube in order to drain itself.
In a multi-pass tube bundle, the lower rows of tubes are severely flooded, because they
have to take the cumulative liquid produced in the tubes of the passes above them in
addition to their own liquid load.

83
It is interesting to note that in all the sulphur condensers used in the sulphur recovery
units, in which sulphur condenses in the tube side, the tilted single pass tube bundle is
used universally to enhance drainage of liquid sulphur out of the tubes. Curiously, the
same consideration seems to be lacking in the application to other tube side condensing
services!

In condenser applications where the vapour to be condensed comes superheated, an


arrangement for desuperheating has to be provided. An example is the superheated
refrigerant vapour from the discharge of a refrigerant compressor, which has to be
condensed totally. Since vapour desuperheating is highly velocity dependent, one is often
forced to use a multi-pass tube arrangement to achieve an acceptable velocity. Also, the
desuperheating heat transfer coefficient is inherently low even at an acceptable vapor
velocity; for the relatively small sensible heat transacted, it requires an inordinate tube
surface area compared to that required for the condensation that ensues in the same
exchanger. The two processes, desuperheating and condensing, have incompatible
requirements: high velocity hence multi-pass arrangement for desuperheating; velocity-
independent hence single pass for condensing to ensure good drainage. Attempting to
realize both desuperheating and condensing in a single exchanger leads to “design
conflict”. How do you get good liquid drainage with a multi pass horizontal bundle? And,
how do you get acceptable velocity with a single tube pass? However, it is quite often
feasible to desuperheat the vapour ex situ. This can be achieved by spraying its own
liquid into the superheated vapour to bring it to saturation before it enters the heat
exchanger. The heat exchanger will then need to be designed for condensing service only.
This reduces the condenser size significantly. For example, in the widely used vapour
compression refrigeration using propane as refrigerant, the compressed propane vapour
can come superheated by as much as 30 °C; if this superheated vapour is saturated ex situ
by spraying liquid propane into the superheated vapour, then the size (tube area) of the
condenser required can be reduced to 50 % of that which is required if the superheated
vapour, without ex situ saturation, is to be desuperheated and condensed in the same heat
exchanger. In the case of an air-cooled condenser, the reduction of the heat exchanger
size (tube area) also means the reduction in the number of fans and fan power. The use
of ex situ saturation for the refrigerant condenser is shown in detail in Section 8.5.2
(Chapter 8: Refrigeration). To be sure, the spraying of liquid into the vapour inlet to the
condenser increases the liquid traffic in the condenser. This increase is not significant:
depending to the degree of superheat in the vapour, the increase is in the order of 5 to 10
weight % of the superheated vapour rate. For a well drained condenser (a tilted single
pass tube bundle, for example), this increase in local liquid traffic in the exchanger is
inconsequential.

5.5 Misconceptions & Disconnects

Many unpleasant surprises arise from the process engineer’s neglect to specify certain
mechanical aspects and geometry that are crucial to the process that is expected to take
place in the heat exchanger. We present the mechanical details of the exchanger that meet

84
the process requirements of a particular service; we also highlight the consequences of
getting these details wrong.

5.5.1 Condensing Services

In condensing services it is important to ensure that adequate liquid drainage be provided


so that the heat transfer surface remains free from flooding. Flooding insulates the tube
surface from the condensing vapour, thus reducing the capacity of the exchanger.
Therefore, in the detailed design of condensing services, every effort must be made to
maximize the heat transfer surface (tube surface) available for direct contact with the
condensing vapour. To minimize flooding for condensing on the tube side a single pass
tilted bundle must be used, especially for total condensation; for condensing on the shell
side of a horizontal exchanger, the vertical baffle cut must be used, with the shell side
outlet nozzle located at the bottom of the shell.

For the mechanical designer focused on the mechanical aspects (which are just as
important as the process requirements), he is concerned with the mechanical integrity of
the exchanger: pressure containment, sealing, tube vibration and the adequacy of support,
etc. He is impartial, for example, whether the baffle cut is vertical or horizontal. He is
interested in baffle spacing for its adequacy in mitigating tube vibration and in providing
structural support to the tube bundle. Usually he is not aware, for example, that baffle cut
orientation is a crucial process requirement unless this information is expressly conveyed
to him. Horizontal or vertical is just six or a half-dozen to him – a vertical cut is just as
mechanically feasible and sound as a horizontal one, and it makes little or no difference
in the mechanical design, fabrication or cost of the exchanger.

In order to express this crucial mechanical requirement and to transmit it unambiguously


every step of the way through detailed design and fabrication, it is a simple matter to
include an entry (or box) in the heat exchanger data sheet by which the designer, in order
to complete the datasheet, is forced to state if he wants the baffle cut to be vertical or
horizontal, or to include an entry in the data sheet to show the angle of tilt for the tube
bundle.

5.5.2 Boiling Services

In boiling services every effort must be made to allow direct contact of the boiling liquid
with tube surface. In the case of boiling on the shell side of a horizontal tube bundle, this
means the tube pitch and pitch angle must be such that they facilitate the dislodging of
vapour from the boiling surface thus minimizing vapour shrouding of the tube area.
Vapour shrouding, also known as vapour logging or vapour locking, deprives the boiling
liquid direct contact with the tube surface, thereby reducing the effective capacity of the
exchanger. For boiling on the tube side, the designer must maintain a two-phase flow
regime that is favorable to boiling. Bubbling flow and annular flow regimes are favorable
to tube side boiling, while mist flow is unfavorable to it. In the bubbling flow regime,

85
which occurs at low vapour to liquid ratio, the liquid remains the continuous phase
carrying along embedded vapour bubbles; the liquid therefore remains in direct contact
with the tube surface, which is a highly desirable situation for boiling. In the annular flow
regime, which occurs at intermediate vapour to liquid ratio, an annular layer of liquid is
sustained inside the tube with the vapour flow in the core of the tube. This regime too
allows direct contact of the boiling liquid with the heat transfer surface. In the mist flow
regime, however, which occurs at high vapour to liquid ratio, the vapour phase becomes
continuous entraining liquid droplets (or mist) in its flow. The continuous vapour phase
effectively insulates the liquid droplets from the heat transfer surface. When the mist flow
regime establishes itself, further boiling stalls and the operating capacity of the exchanger
is thereby severely curtailed.

Another important consideration in the design of an exchanger for boiling services is the
temperature of the heating medium itself. There is a limit in the temperature difference
between the hot side and the boiling (cold) side. If this temperature difference is too
great, a relatively stable layer of vapour forms, hugging the tube surface on the boiling
side. This shrouding prevents the boiling liquid from direct contact with the tube surface,
similar to the vapour logging phenomenon mentioned above. Usually the limit of
temperature difference is expressed, indirectly, as a heat flux limit. This heat flux is the
rate of heat transfer per unit tube surface area, given the temperature difference (between
the hot and cold fluid) and assuming normal boiling coefficient, normal boiling
coefficient being one that obtains at bubbling or annular flow regime. By conventional
wisdom the heat flux limit is about103 kW/m2 [10,000 Btu/hr ft2]. A heat exchanger
designed for a boiling service without regard to the flow regime of the boiling liquid or to
the heat flux limit will fall far short of the intended capacity, often dramatically and
“mysteriously”.

5.5.3 NTIW Tube Bundle

The tubes in window are those tubes that have a span equal to twice the baffle spacing, as
shown in Figure 5.5 (a). When these tubes in window are removed, the resultant bundle is
said to have an NTIW (no tubes in window) geometry. The NTIW geometry is used when
the tube span in window is too great to prevent tube vibration. However, there are
applications where the NTIW geometry is required by process consideration alone,
having nothing to do with mitigating tube vibration; in other words, this geometry is
imposed by the process at hand, whether or not the tube span in window happens to be
adequate for preventing tube vibration. An important example of these applications is a
reboiler with boiling taking place in the shell side of a horizontal exchanger. In addition
to NTIW, the boiling process also requires the following additional mechanical details;
see Figure 5.6:

• The parts of the baffles protruding into the top and bottom window be truncated
(removed) to allow the shell spaces at the top and bottom of the tube bundle to act
as header spaces for shell side outlet and inlet, respectively;

86
• Baffle spacing be determined by mechanical considerations only (tube vibration,
tube support), since it has no bearing on the process performance, thermal or
hydraulic;
• TEMA X for the shell side flow pattern (boiling liquid) be used for thermal and
hydraulic calculations (simulation). The upward cross flow pattern is the natural
consequence of the buoyancy created by the boiling liquid. The bottom and top
shell spaces, i.e. the NTIW spaces, act as headers to facilitate inlet distribution
and outlet collection.

Top of shell Span in w indow


A = 2 x baffle spacing
Tubes
in w indow

A
Bafffle spacing View A-A
(a) TEMA E shell with full tube bundle; horizontal baffle cut Full bundle

Top of shell Span


B =1 x baffle spacing No tubes
in w indow
(NTIW)

B
Bafffle spacing
View B-B
(b) TEMA E shell with NTIW bundle; horizontal baffle cut NTIW bundle

Figure 5.5:
Full bundle and no-tubes-in-window (NTIW) bundle

Therefore the process designer needs to convey these requirements clearly to the
mechanical designer and, by extension, to the fabrication shop. In particular, he should
stress that baffle spacing has no bearing on thermal and hydraulic performance, leaving
the mechanical designer the freedom to determine this spacing from mechanical
considerations alone without any process constraints. He should also specify the need to
truncate the baffle protrusions to create unobstructed header spaces for inlet liquid
distribution and outlet flow (two-phase) collection. He should note that these baffles are
“non-standard” in that they are neither single segmental nor double segmental.(See
Figures 5.1 and 5.2.) Furthermore, he should specify that supports against the bending
moment of the tube bundle are to be non-obstructive to longitudinal flow in the lower
header space in the shell. All these crucial mechanical details can be articulated
unambiguously as notes and sketches forming part of the data sheets for transmission to
detailed design and fabrication of the heat exchanger.

A typical physical inlet and outlet arrangement is shown in Figure 5.6. It is noted that
with the adequate top header space provided, only one outlet nozzle may be adequate.
This greatly simplifies the outlet piping, an important consideration especially for the

87
thermosiphon reboiler; with a single outlet nozzle less elbow room is required of the
outlet piping; this in turn results less elevation difference between the thermosiphon
reboiler and the fractionation tower.

Heat Liquid+vapour Liquid+vapour Header


medium A space
Baffle spacing

Baffle w ith minimum


Heat A Header
Liquid protrusion into NTIW
medium space
View A-A:
spaces

The prevailing flow patern of the boiling liquid is cross flow upw ard (flow pattern
contemplated for a TEMA X shell ). Adequate header space must be provide above and
below the NTIW bundle to facilitate this natural cross flow pattern. Not show n, for the
sake of clarity, are supports against the bending moment of the bundle. These supports
must be unobstructive to flow in the bottom header space.

Figure 5.6:
TEMA X shell for shell side side boiling; adequate header spaces above and below tube bundle

5.5.4 Shell Side Condensing

It is obvious why a shell side condenser cannot function with the horizontal baffle cut.
See Figure 5.7. Severe flooding occurs and the condensing capacity is lost the moment
one attempts to put the condenser into operation. As liquid accumulation at the bottom of
the shell reduces the condensing surface and seals off the vapour flow path, pressure is
backed up until the liquid seal is broken by the pressure back-up. But liquid drop-out
continues to replenish this liquid seal. This erratic pressure swing usually makes the unit,
of which this exchanger is a part, inoperable.

In one instance, the exchanger with the wrong (i.e. horizontal) cut orientation was
intended to be the overhead condenser of a depropanizer tower. Over-pressure in the
tower occurred seconds after heat to the reboiler of the tower was turned on. The
condenser simply refused to condense the vapour generated by the reboiler! Intense
trouble-shooting activities ensued, which included removing the tube bundle for thorough
examination of the condenser inside and out. Thorough examination was also conducted
of the design calculations and detailed mechanical drawings. Nothing untoward was
found. The exchanger was put back together and several more futile attempts were made
to start up the depropanizer. The analysis, diagnostics, prognostics and technical debate
which were proffered during the repeated and long trouble-shooting exercises are rather
amusing in retrospect. But nothing was amusing at the time. It took the casual
observation of a totally “disinterested” individual who exclaimed - in all innocence

88
reminiscent of the boy in H. C. Andersen’s The Emperor’s New Clothes - , “but your
baffle cut is horizontal!”

Coolant
Vapour Top of shell
B Horizontal baffle cut

B
Coolant Liquid build-up Expected
condensate (b) View B-B

(a) Elevation: TEMA E shell w ith horizontal baffle cut;


liquid build-up at the bottom of shell seals
off vapour flow path.

Figure 5.7:
A condenser with horizontal baffle cut; designed to fail

Fortunately, it was a relatively simple exchanger. The bundle was pulled out of its shell,
rotated 90 degree and re-installed. External modification of the cooling water piping to
accommodate this re-oriented bundle was also easily accomplished. The exchanger
functioned as intended thereafter

Baffle cut orientation may be of no consequence to the mechanical designer, but it can be
vital to the process. The whole unit can be rendered inoperable by one design flaw:
wrong baffle cut orientation.

5.5.5 Abuse of the TEMA J Shell for Boiling Service

It cannot be over-emphasized that the flow pattern of boiling liquid in the shell side using
a horizontal bundle is perforce cross flow, i.e. flow across the bundle as contemplated by
the TEMA X shell, because boiling liquid flows upward as bubbles are formed. The
designer has no choice in this; he has to use a TEMA X shell for this application of shell
side boiling, as shown in Figure 5.6. Failure to recognize this perforce cross flow
phenomenon often results in designs that impede or make impossible this flow pattern
which has to prevail in shell side boiling.

This obligatory use of the TEMA X shell for shell side boiling is in contrast to
applications where the shell side fluid is single-phase (either all liquid or all vapour) and
suffers no phase change during the process; here the choice of the TEMA X shell may be
made either to circumvent a temperature pinch in order to eliminate the need to use multi-
shells in series, or to stay within the allowable shell side pressure drop; that is to say, the
choice of TEMA X shell is made for the advantages it has over the other TEMA shell
types in the context of the particular set of parameters (temperature profile; allowable
pressure drop) imposed by the application in question. In other words, given a different

89
set of process parameters, TEMA X may or may not be a good choice in single phase
service against other TEMA shell types.

The TEMA J shell divides one inlet entering at the middle of the shell into two outlets
leaving near the ends of the shell. In single phase service the inlet can be located at the
top of the shell, with outlets at the bottom or vice versa, as shown in Figure 5.8.

Shell side
Shell side
outlet Shell side
outlet Top of shell Top of shell
inlet

Shell side Shell side Shell side


inlet outlet outlet
(a) Bottom inlet (b) Top inlet
Figure 5.8:
TEMA J shell; or shell of divided flow

J Shell with Vertical Baffle Cut

The TEMA J shell has been used in the design of many horizontal reboilers, where the
liquid enters at the bottom and the two phase flow leaves at the top of the shell. One
possibility of the design is shown in Figure 5.9, where we have a full bundle with vertical
baffle cut. The TEMA J shell contemplates a flow pattern depicted in Figure 5.9 (c),
whereby the whole fluid meanders from one baffle space to the next, much like the
movement of a snake on a horizontal plane. This flow pattern can be sustained as long as
the fluid remains single-phase. For boiling in a horizontal shell, however, this serpent-
like meandering flow pattern is not sustainable. The flow is perforce cross flow by the
very nature of boiling, as shown in Figure 5.9 (a). The thermal and hydraulic calculations
(simulation) for boiling based on the TEMA J shell flow pattern are therefore entirely
fictitious. Compared to the actual cross flow pattern, which is the pattern contemplated
by the TEMA X shell, both the shell side heat transfer coefficient and the pressure drop
calculated from this fictitious pattern (presumed to be sustainable in the TEMA J shell)
are over-estimated (i.e. higher than they really are). The over-estimate of the heat transfer
coefficient means that the exchanger is undersized, because in operation the cross flow
pattern in the manner of the TEMA X shell will prevail – unintended and unbeknownst to
the designer. If the exchanger is meant to be a thermosiphon reboiler, the over-estimate of
shell side pressure will result in the designer specifying an unduly large elevation
difference between the fractionation tower and the reboiler and have significant impact
on the infrastructures and other ancillary equipment of the tower and hence will
needlessly raise the installed cost of the tower.

90
Heat Liquid+vapour Liquid+vapour
medium Top of shell
A

B
A B
Heat
Liquid (b) View A-A
medium
(a) Elevation: Flow pattern in operation is perforce cross flow

(c) View B-B (plan view ): Intended flow pattern of total fluid (Liquid + vapour)
meandering serpent-like from one baffle space to
the next is fictitious, w hen boiling occurs.

Figure 5.9:
TEMA J shell intended for shell side boiling - vertical baffle cut, full bundle

Quite apart from the fictitiousness of calculations, the TEMA J shell is not conducive to
the cross flow pattern, even though this cross flow pattern will exert itself in actual
operation, regardless of shell side geometry. In the TEMA J shell, there is no header
space (freeway) for the inlet to spread longitudinally along the bottom of the tube bundle;
and likewise no header space over the top of the tube bundle for the collection and
channeling of the outlet. These spaces are obstructed by tubes and baffles. Because of the
lack of freeway on the top of the bundle, vapour tends to break out and hold up at the top
of the bundle. The vapour holdup effectively shields the top rows of tubes, making their
surface unavailable for boiling. This diminishes substantially the operating capacity of
the reboiler.

Some designers recognize the detrimental effect of vapour holdup on operating capacity.
To circumvent this a truncated bundle is used, which has no tubes in the top and bottom
rows, to create top and bottom header spaces; but otherwise the design is still that of a
TEMA J shell with vertical baffle cut. (See Figure 5.10.) (Note that a bundle so truncated
is not the same as the NTIW bundle for TEMA X shell, which has no baffle protruding in
the window. (See Figure 5.6.) The thermal and hydraulic calculations are still based on
TEMA J flow pattern, even although this flow pattern cannot be sustained due to boiling.
The baffle protrusions of into the header spaces still represent a considerable obstruction,
especially for the collection and longitudinal routing of the two-phase fluid (product of
boiling) to their outlet nozzles.

It should be noted that the full-bundle J shell flow pattern may be well suited to
applications where there is no phase change of the fluid traversing the shell side. In other
words, there is nothing wrong with J shell provided it is appropriate for the service.

91
Heat Liquid+vapour Liquid+vapour
medium Top of shell
A No tubes
(header spaces)

Heat A
medium Liquid

(a) Elevation: Actual cross flow pattern (b) View A-A

Figure 5.10:
TEMA J shell intended for shell side boiling - vertical baffle cut; bundle trucated to
make header spaces at top and bottom; protrusions of baffles into header spaces
still present obstructions to flow longitudinally

J Shell with Horizontal Baffle Cut

When looking at a shell side boiling design, for example in a horizontal thermosiphon
reboiler for a fractionation tower, many designers would instruct the program to size and
design it as a TEMA J shell type. Figure 5.11 shows the J shell flow pattern with a
horizontal baffle cut. Based on the baffle cut and spacing specified by the designer, the
program will indeed proceed to size and design it as a TEMA J shell. Heat transfer and
hydraulic calculations are performed, as well as the mechanical evaluations are made,
based on this J flow pattern, i.e. the serpentine meandering a mixture of liquid and vapour
from one baffle space to the next.

The flow pattern as contemplated for the J shell with horizontal baffle cut is just as
fictitious when applied to boiling as that of a TEMA J shell with vertical baffle cut
discussed earlier. The boiling liquid plus the vapour engendered can’t possibly meander
upward and downward from one baffle space to the next to reach the outlet nozzles at
both ends of the shell. Instead, it rises upward, as it must due to the buoyancy created by
boiling; and as it reaches the top part of the shell it disengages the vapour. As shown in
Figure 5.11 (c), the disengaged vapour is trapped in the upper baffle spaces. The trapped
vapour, replenished by boiling, tends to grow in volume and, as it does, the liquid flow
path is hindered. Liquid is forced to flow through a small gap (between the liquid-vapour
interface and the horizontal edge of a baffle in its attempt to reach the next baffle space.
With the small gap available as flow area, the liquid flow at a high velocity entraining
some vapour into its flow. The liquid-vapour interface fluctuates as a result of the tug-o-
war between the tendency of the trapped vapour to grow and liquid forcing its flow at
high enough velocity to entrain vapour to limit its growth. This is a bad situation, which
manifests as a high and fluctuating pressure drop. The high pressure drop can easily
exceed the margin of the head difference required to drive a thermosiphon reboiler. The
trapped vapour, shrouding the tubes in it, reduces the heat transfer area available for
boiling. Also, in half of the baffle spaces the liquid flows downward. This downward

92
flow of liquid runs against the rising bubbles of vapour and results in a high vapour
holdup in those baffle spaces. The thermosiphon reboiler either stalls due to insufficient
head to drive itself; or, when flow does establish, albeit erratically, not enough vapour is
generated due to vapour shrouding of tube surface. A fractionation tower connected to the
reboiler will fail dramatically by weeping its trays due to low vapour flow to the trays.

Heat Liquid+vapour Liquid+vapour


medium Top of shell
A

Heat A
medium Liquid (b) View A-A

(a) Elevation: TEMA J shell w ith horizontal baffle cut; intended flow
pattern of total fluid meandering up and dow n baffle
spaces is not obtainable for boiling service

Trapped vapour
Vapour-liquid interface, erratic
and restricting liquid flow
Bubbles of vapour Dow nw ard flow of liquid against rising
from boiling vapour creates large vapour holdup
and vapour shrouding of tube surface
(c) Trapped vapour tending to seal off liquid flow path

Figure 5.11:
A reboiler designed to fail

Yes, shell side reboilers have been installed with horizontal baffle cuts! The proud
owners of these monstrosities are usually left holding the bag while entertaining both the
comedy and the tragedy of the situation.

5.6 Over-reliance on Flagging or Warning

What follows is not a review or critique of any commercial software used for shell and
tube heat exchanger design and simulation. The comments here are made on the
unwarranted and often mistaken expectations on the part of the software user. The over-
reliance on the software to discover a user’s mistakes and to warn him about them has
been a source of many unpleasant surprises. It must be reiterated here that this in no way
reflects on the quality of any commercial software, most of which (the successful ones)
have good data base, good correlations, sound calculation principles and a robust
numerical process. It is recognized that flagging or warning that is currently unavailable
in a software can be easily implemented, if one so desires. The software developer is
usually quite amenable to implementing user-suggested features to enhance the
usefulness and to gain wider acceptance of his software. For example, the software can

93
easily be programmed to give a warning message, such as: Boiling in shell side detected;
TEMA J shell flow pattern as specified cannot be obtained; calculations aborted; please
consider other options; or, Phase change detected on shell side; incompatible with the
horizontal baffle cut specified; run aborted; please consider other options.
Implementation of this sort is not a question of feasibility but one of choice, a collective
choice of the software users and software developers. The choice is: to what extent, in
our design effort, do we want to relinquish to computer programs our thinking,
circumspection, intuition and understanding? Design is an intimately human enterprise;
the computer programs are just a tool. It may be recalled not long ago the only tools for
numerical work at designers’ disposal were the slide rule and mathematical tables – the
hand-cranked adding machine being a luxury. This is no nostalgic musing, but a tool is a
just tool and is never a substitute for careful thinking.

5.6.1 Warning on Exchanger Geometry

The popular commercial software for the design and simulation of the shell and tube heat
exchanger is programmed with many useful warnings, which are posted in the program
output. Among these warnings are those which indicate, for example, the possibility of
tube vibration due to inappropriate tube span; excessive velocity in certain part of the
exchanger; erosive flow conditions; flow regimes that may induce vibration, etc. What
does it warn us against, and what does it not? Caveat emptor: the onus is on the user to
find out. Pertinent to the exchanger geometry are questions such as: does it warn against
the following?

• Use of a tube bundle with horizontal baffle cut for condensing in a horizontal
shell;

• Use of a tube bundle with horizontal baffle cut for two-phase flow in a horizontal
shell;

• Use of a tube bundle with horizontal baffle cut for boiling in horizontal shell;

• Use of the TEMA J shell for boiling in a horizontal shell;

• Use of a TEMA shell type other than TEMA X for a boiling in a horizontal shell;

• Insufficient header spaces (NTIW spaces), or none, provided for boiling in a


horizontal TEMA X shell (see Figure 5.6); and

• Use of wrong nozzle locations – for example, placing a tube side nozzle on the
wrong quadrant of the channel head for a multi-pass tube bundle.

In the years when the present author was actively engaged in process and equipment
design (2010 and earlier), the answers to all the above questions were no. (As pointed out
earlier, it is quite easy to implement the flagging for the above incidents of wrong

94
geometry; it is only a question of choice and desirability.) It is important to be aware of
this, so that we don’t defend a wrong design by saying that the program does not warn
against us it, as some designers are wont to do. Such defense is of course untenable.

5.6.2 Manuals, Texts and Literature

Some manuals and texts present the TEMA J shell implicitly as the one to be used for a
horizontal reboiler - implicitly, because the horizontal reboiler in its inlet and outlet
piping scheme resembles the symbolic representation of the J shell in the TEMA guide on
shell types. The icon of a horizontal reboiler is shown, almost invariably, with the liquid
entering the bottom of the shell at one point and the vapour-liquid mix leaving from two
points at the top of the shell. (See Figure 5.8 (a).) This is a powerful imagery that inspires
and even appears to compel the designer to adopt the TEMA J shell for his horizontal
reboiler. This is unfortunate, as we have seen the TEMA J shell with horizontal baffle
cut can’t possibly function as a reboiler or, for that matter, in any application with shell
side boiling or condensing. The TEMA J shell with vertical baffle cut stands a chance, to
be sure; but it comes with a heavy penalty due to the loss of heat transfer surface, a fact
that often eludes the designer. (See Section 5.5.5.) The program, as we have noted,
proceeds with the thermal and hydraulic calculations as if the TEMA J shell flow pattern
could indeed be sustained, without warning the user that it is in fact physically infeasible.

5.6.3 TEMA X Shell for Boiling

As we have seen boiling liquid rises upward across the entire length of the horizontal
tube bundle. The suitable shell side geometry should be one that facilitates this
phenomenon. This calls for a truncated tube bundle (NTIW and truncated baffles) that
provides an unobstructed header space below the bundle for the inlet liquid to distribute
with ease in directions longitudinal to the bundle, as well as an unobstructed header space
above the bundle for the boiling product (vapour or a two-phase fluid) to flow
longitudinally to its outlet nozzles. (See Figure 5.6.) The vapour break-out in this upper
header space does not result in vapour shrouding of the tube bundle, since there are no
tubes in the header spaces. Also, more realistic thermal and hydraulic calculations are
made, now that the program is instructed to proceed with these calculations on the basis
of the flow pattern contemplated by the TEMA X shell, namely the flow pattern that is
imposed by the nature of boiling.

It should be noted that the TEMA X shell pertains only to the flow pattern to be used for
thermal and hydraulic calculations; it does not imply or contemplate any specific tube
bundle geometry. When the TEMA X shell is specified as input to the design software,
thermal and hydraulic calculations (film coefficients, pressure drop, etc) are made on the
basis that the total shell side inlet fluid is distributed evenly over the whole length of the
tube bundle; the inlet fluid so distributed then flows across the bundle, perpendicular to
the tube length. The flow pattern can be obtained physically provided the exchanger has
the appropriate tube bundle geometry. The desired geometry mentioned above (Figure

95
5.6) for the horizontal reboiler must be explicitly indicated by the process designer on the
heat exchanger data sheet and passed on as the basis for detailed mechanical design and
fabrication. Indicating that the exchanger shell is TEMA X says nothing to the
mechanical designer about the tube bundle. In particular, it must be shown in the data
sheet how the tube bundle and baffles are to be truncated to obtain the required top and
bottom header spaces. No warning is given by the heat exchanger design software, for
example, if these header spaces are not provided. Without these header spaces, the inlet
fluid cannot physically distribute evenly over the length of the bundle; the vapour
generated from boiling has no free passage to the outlet nozzle(s); it gets trapped in the
upper half shell and shrouds the tube area. The software does not recognize this physical
infeasibility brought about by the absence of appropriate header spaces; instead, it
proceeds “blissfully” with the cross-flow thermal and hydraulic calculations as
contemplated by TEMA X shell and yield its results to the user.

5.6.4 Nozzle Size and Orientation

Heat exchanger design software may require the user to state the nozzle size and
orientation for the inlet and outlet nozzles on the shell side and tube side of the
exchanger. The nozzle size is used in the design calculations in the software to ensure
that the limit on the velocity of the fluid passing through the nozzle stays below a certain
limit. This limit is expressed by the quantity ρu2, where ρ is the fluid density and u is the
velocity of the fluid passing through the nozzle. The conventional limits are: ρu2 be equal
or less than 8925 kg/m s2 [6,000 lb/ft-sec2] for tube side nozzles and be equal or less than
5950 kg/m s2 [4000 lb/ft-sec2] for shell side nozzles. Nozzle orientation, on the other hand,
is handled by the software only as an instruction, presented in the printout in the form of
a sketch, to be passed on to the mechanical designer and, eventually, to the fabricator.
Nozzle orientation plays no part in the thermal and hydraulic calculations performed by
the program. For example, if one has an even number of baffle intervals (i.e. even
number of cross passes in the shell side) in a TEMA E shell, one must have the inlet and
outlet nozzles on the same side of the shell. Or, as another example, in a condensing
service in the horizontal shell side, the shell side outlet nozzle must be placed at the
bottom of shell. Any misplacement of nozzle made by the designer, whether on the shell
side or the tube side, will not be flagged in the program printout; and, more often than
not, this misplacement will go unnoticed in spite of all the rigour in multidisciplinary
design review, laying the groundwork for yet another instance of an unpleasant surprise
in the field.

5.7 Temperature Profile – Countercurrent and Cocurrent

The required area of a heat exchanger is inversely proportional to the log mean
temperature difference (LMTD) between the shell and tube side: a higher LMTD means
more heat transfer per unit of exchanger area. As can be demonstrated in a shell and tube

96
exchanger with a single tube pass, the countercurrent arrangement affords a more
favorable LMTD. See example in Figure 5.12, where the LMTD is given by:

T −t 
LMTD = [(T1 − t 2 ) − (T2 − t1 )] ÷ ln 1 2  , countercurrent arrangement (5.1)
 T2 − t1 
 T −t 
LMTD = [(T1 − t1 ) − (T2 − t 2 )] ÷ ln 1 1  , cocurrent arrangement (5.2)
 T2 − t 2 

T1 = 100 oC

T2 = 80 oC

Temperture
Hot inlet, T1
t2 = 50 oC LMTD = 47.5 oC

Cold inlet, t1 t1 = 35 oC
Cold outlet, t2
Hot outlet T2
Heat transaction
Countercurrent arrangement Countercurrent temperature profile

T1 = 100 oC

Hot inlet, T1 T2 = 80 oC
Temperture

t2 = 50 oC

t1 = 35 oC
Cold outlet, t2
LMTD = 45.3 oC
Cold inlet, t1
Hot outlet T2 Heat transaction

Cocurrent arrangement Cocurrent temperature profile

Figure 5.12:
Countercurrent and cocurrent arrangements:
Temperature profile and LMTD

It is clear that cocurrent arrangement is less effective than the countercurrent due to the
debit of LMTD alone. However, the disadvantage of the cocurrent arrangement becomes
much more remarkable when the outlet temperature of cold fluid is intended to be higher
than the outlet temperature of the hot fluid as a process requirement. With the
countercurrent arrangement, this process requirement can be met; but not with the
cocurrent arrangement. An example is given in Figure 5.13, where two identical
exchangers (each with one shell pass and one tube pass) are shown in the countercurrent
and cocurrent arrangements. It is seen that the process requirement is met with the
countercurrent arrangement, whereas with the cocurrent arrangement there is a significant
(21 %) shortfall in duty and missing of temperature targets. See Appendix VI for the
calculations of the pinch temperature and duty shortfall.

97
T1 = 75 oC

o
LMTD = 16.6 C

Temperture
Hot inlet, T1

75 C
o T2 = 33 oC, target met
t 2 = 45 oC,
Cold inlet, t1 target met
o
25 C
Intended duty achieved t1 = 25 oC
Cold outlet, t 2 Hot outlet, T2
o o
45 C 33 C
Heat transacted
Countercurrent arrangement
Countercurrent temperature profile and LMTD

o
T1 = 75 C o
LMTD = 13.1 C
Hot inlet, T1
o
o
75 C Temperture T2C= 41.9 C, target missed
Cold outlet, t C T pinch = 41.1 oC
o 2
40.7 C
Cold inlet, t1 Hot outlet, T2C t 2C o
= 40.7 C, target missed
25 C
o o
41.9 C
t1 = 25 o
C
Only 79% of
intended duty
Cocurrent arrangement
Heat transacted
Cocurrent temperature profile and LMTD
Figure 5.13 :
Cocurrent pinch causing a shortfall in duty

In Figure 5.13 the temperature profiles of the fluids are plotted against heat transacted. It
is instructive to plot these temperature profiles along the tube length, as shown in Figure
5.14. In the cocurrent arrangement, the hot fluid temperature drops rather rapidly after
entry (compared to the countercurrent arrangement), with an equal rapid rise of the cold
fluid temperature. After the rapid changes, both temperatures approach the pinch
temperature asymptotically. (See Figure 5.14 (b).) The hot fluid exits the exchanger at a
temperature just higher than the pinch temperature, while cold fluid exits it at a
temperature just below the pinch. When asymptotic temperature profiles such as those
just described are obtained in an exchanger, it is customary to say a “temperature pinch”
occurs or exists in that exchanger. But it is rather imprecise to say that because the pinch
temperature is approached but not quite reached inside or at the exit of the exchanger.
(For both the hot and cold temperatures to actually reach the pinch, an infinite tube area
would be required.) Imprecise as it may be, we shall adhere to the customary description
in the narrative of this chapter while mindful that so-called pinch temperature is never
actually reached at any point on the tube surface. Note that in the countercurrent
arrangement, the temperature profiles exhibit no such asymptotic behavior. (See Figure
5.14 (a).) In that case we say that temperature pinch does not occur with the
countercurrent arrangement.

98
T pinch = 41.1 oC
o o
T1 = 75 C T1 = 75 C (Asymptote)
Temperature

Temperature
o
T2C = 41.9 C
T2 = 33 oC
t 2 = 45 oC t1 = 25 oC t1 = 25 oC t 2C = 40.7 oC

Inlet/outlet Outlet/inlet Inlet Outlet


Tube length Tube length

(a) Countercurrent (b) Cocurrent

Figure 5.14:
Temperature profiles versus tube length

5.7.1 Cocurrent Passes in a Shell and Tube exchanger

In a shell and tube exchanger with multiple tube passes and a single pass in the shell side,
half the tubes are in countercurrent flow and the other half in cocurrent flow with respect
to the shell side flow. See Figure 5.15.

Cold outlet, t2
o
45 C (intended) o LMTD correction factor = 0.01
o Hot inlet, T1 75 C
<45 C (actual)

o
Pinch (about 40 C) Hot outlet, T2
Cold inlet, t1 o
o occurs somew here in 33 C (intended)
25 C this cocurrent pass o
>33 C (actual)
Figure 5.15:
Two tube-pass and single shell-pass exchanger:
(For clarity, only one U tube of the bundle is shown; shell side baffles are not shown)
The upper tubes are countercurrent and the lower ones cocurrent.
A temperature pinch occurs somewhere in the cocurrent path since t 2 > T2 as intended.
The intended targets t 2 and T2 will not be met and there will be shortfall of duty.

A shell and tube exchanger with one shell pass has the TEMA E shell type. If the
intended hot outlet temperature is lower than the cold fluid outlet temperature a
temperature pinch will occur somewhere along the one of the cocurrent passes and
cooling of hot fluid down to and below the pinch temperature is impossible. Quite apart

99
from the possibility of a temperature pinch, a cocurrent path also has an inherently lower
LMTD for the following reason: the shell side baffles cause the shell side flow to
meander from one baffle space to the next giving rise to a predominantly cross flow in
the baffle spaces. The cross flow is a deviation from the true countercurrent flow and is a
debit against the effective LMTD.

Because of all these debits brought about by the multi-pass tube bundle in a shell that has
a single pass, the exchanger loses effectiveness. This loss in effectiveness is expressed as
a correction factor to be applied to the LMTD calculated from the intended inlet and
outlet temperature of the two fluids, as given by (5.1). For a perfect countercurrent flow
this correction factor is unity. For a multi-pass tube bundle, it is less than 1. It should be
noted it is less than unity even in the absence of a temperature pinch because there is
always a loss of effectiveness due to (a) the inherently lower LMTD along the cocurrent
path of a multi-pass tube bundle; and (b) the deviation from the true countercurrent flow
in the shell side, where cross flow in the shell side is induced in the baffle spaces. The
lower the hot fluid outlet temperature than the cold fluid outlet temperature the lower this
correction factor becomes.

As can be seen from a sample LMTD correction chart in Figure 5.16, any correction
factor below 0.85 is unacceptable because of very sharp drop of the curve around 0.85.

This means a correction factor of 0.70, for example, calculated from the intended inlet
and outlet temperatures could very well be 0.25 in operation (instead of 0.70), because of

100
the steep drop of the curves in the correction chart. This steepness means uncertainty; it
means there is practically no distinction (or resolution) between 0.70 and 0.25 along the
steep drop. In practice, therefore, a minimum acceptable correction factor 0.85 should be
adopted in order to stay well clear of the steep part of the curve so that the correction
factor won’t “fall off the cliff”, so to speak. When an operating exchanger’s correction
factor falls off the cliff, heat transfer just stalls. This means that temperature pinch occurs
in a cocurrent pass much sooner (or much closer to the hot fluid inlet in terms of flow
path length) than implied by the calculated correction factor of 0.70. It may be hard to
imagine how a shell and tube exchanger, passive, with no moving parts whatsoever and
with no external motive power to drive any mechanism in it, can stall to a point to force a
plant shutdown. It can and it does!

Empirical equations and charts are available to evaluate this correction factor; these are
illustrated in Appendix VII for the example shown in Figure 5.15. It is seen there that for
this example an LMTD correction factor of 0.01 is obtained. This extremely low
correction factor of 0.01 means that it has “fallen off the cliff”, as well as fallen off the
scale of the chart. It is an indication that a temperature pinch exists in the exchanger; and
in order to meet process targets the temperature pinch must be circumvented by a revised
exchanger design.

5.7.2 Circumventing Temperature Pinch by Providing More Shell Passes

When a pinch occurs in a design of a multi-pass tube bundle in a single shell, as


manifested by a low, “fallen-off-the-cliff” LMTD correction factor, the number of shell
passes must be increased to circumvent the pinch.

The number of shell passes can be increased by:

• Putting two or more single-pass shells (TEMA E shells) in series; or

• Using a longitudinal baffle in the shell to create two passes in a single shell. A
shell with a longitudinal baffle to create two passes in it is designated TEMA F
shell type.

Consider the process requirement of the example shown in Figure 5.15 above. If we
configure two TEMA E shells in series, which increase the shell side passes from one to
two, the pinch is circumvented; and process targets (duty and temperatures) will be met.
Note that in this two-shell in series configuration (Figure 5.17), we still have half of the
total tube passes in cocurrent, but the fluid temperatures along these cocurrent paths no
longer result in a pinch. As shown in Appendix VII, the LMTD correction factor is
increased from 0.01 to an acceptable 0.853 when two TEMA E shells are configured in
series.

101
Cold outlet, t2
o Hot inlet, T1
45
o LMTD correction factor = 0.853
(target met) 75 C

Cold inlet, t1
Hot outlet, T2
o o
25 C 33 C (target met)

Figure 5.17:
Two TEMA E shells in series to circumvent temperature pinch.
(For clarity, only one U tube is shown in each shell; and shell side baffles are not shown)
Process temperature targets are met.

Another way to increase the number of shell side passes is to use the TEMA F shell type
as shown in Figure 5.18.

Cold outlet, t2
o
45 C o
Hot inlet, T1 75 C
(target met)

Cold inlet, t1 Hot outlet, T2 Longitudinal baffle


o o
25 C 33 C (target met)

Figure 5.18:
TEMA F shell type:
Two shell side passes in a single shell accomplished by the use of a
longitudinal baffle.(For clarity, the segmental transverse baffles in the
shell side are not shown)

It is noteworthy that in both the two tube passes in the single TEMA F shell the flow is
countercurrent. (See Figure 5.18). This is in contrast to the use of two TEMA E shells in
series, because in this case, half the tubes in each of the two shells are still cocurrent.
Without the presence of cocurrent passes, which have inherently lower effective LMTD,
one should expect the tube area required by the one TEMA F shell as shown in Figure
5.18 to be somewhat less than the total tube area in the two TEMA E shells connected in
series as shown in Figure 5.17. However, this advantage of increasing number of passes
in the shell side by the use of a single TEMA F shell, versus the use of two TEMA E

102
shells, is not distinguished in published empirical formulas or charts for LMTD
correction. It is not clear, either, if this advantage of the absence of cocurrent paths in the
TEMA F shell is accounted for in commercial software for the shell and tube exchanger
design. One can surmise, therefore, that the tube area in the TEMA F shell is somewhat
over-estimated, thus allowing one to err on the conservative side. This slight over-
estimate of tube area notwithstanding, the obvious, tangible cost advantage of using a
single TEMA F, versus two TEMA E shells, is that one gets to circumvent temperature
pinch in one single physical shell (instead of two shells in series).

5.7.3 Eliminating Pinch by Changing Process Temperatures

It is seen that in a single-pass shell if the effect of a temperature pinch is unacceptable


(low LMTD correction factor), the pinch can be eliminated by:

• Lowering the cold fluid outlet temperature (by increasing the flow rate of the
coolant, for example);

• Raising the hot fluid outlet temperature; or

• Doing both of the above.

However, one must carefully consider the ramification of making such temperature
changes and its impact on the rest of the process scheme. Heat exchange between a hot
and cold process fluid may be, for example, part of the heat recovery measure involving a
train of many inter-related exchangers. Such temperature changes are sure to require the
revision of the design of the whole train. Or as another example, the hot fluid with its
raised outlet temperature may have to be cooled further to its original intended outlet
temperature, because this intended temperature is an important process parameter. This
further cooling then requires additional exchangers. Therefore, one usually fares better to
circumvent temperature pinch by increasing shell passes by either connecting two TEMA
E shells in series or using TEMA F shells, than to vary any process stream temperature.

It is observed that, when the cold fluid in question is cooling water, the designer is
tempted to take the liberty, unilaterally and unbeknownst to his client, to use a lower
cooling water outlet temperature to circumvent the temperature pinch in order to stay
with one shell with single shell pass, i.e. one single exchanger with TEMA E type shell.
One could surmise that such a penchant to lower cooling water outlet temperature arises
from his unawareness of the impact of his action on the entire cooling water system. But
quite often, lowering cooling water outlet temperature is done intentionally to short-
change his client on equipment sizes in a fixed price contract. The lowering of cooling
water outlet temperature results in a significant increase in cooling water demand for the
exchanger in question. If this apparently innocuous unilateral relaxation (lowering) of
cooling water return temperature is widely practiced, the cooling water system can
quickly run out of capacity. The consequences of lowering or raising cooling water outlet
temperature are discussed in Section 5.7.4.

103
5.7.4 Abuse of the Cooling Water Return Temperature

The cooling water system consists of cooling towers, circulating pumps, distribution (or
supply) and collection (or return) headers and laterals connecting these headers to the
individual operating units. The use of cooling water is justified over the air-cooled
exchangers for larger installations where the economy of scale prevails over the use of a
large number of air-cooled exchangers (also known as fin-fan coolers).

Cycle of Concentration

The cooling water outlets from all the water-cooled exchangers in the operating units of
the complex are collected into the return header. The return header carries this warm
cooling water to the cooling tower where it is cooled and recirculated by being pumped
into the supply header to distribute back to the users. Cooling is achieved in the cooling
tower by contacting the water, cascading down from the top of the tower to the retention
basin at the base, by a cross current of ambient air flow induced by fans installed on the
top of the tower. Two mechanisms are at play in this cooling process: sensible heat
transfer from water to air and latent heat of evaporation. Besides cooling the water,
evaporation has two effects: loss of cooling water to the atmosphere and concentration of
the total dissolved solid (TDS) in the inventory of water held in the circulating cooling
water system. To maintain the TDS at an optimally tolerable level, water from the
inventory must be continuously purged (or blown down). This blowdown and the
evaporative loss require that fresh make-up water be continuously added to the cooling
tower water basin to maintain the inventory level. The cycle of concentration (COC) is
the ratio of the TDS maintained in the circulating cooling water to the TDS in the fresh
make-up water.

The higher the COC at which the cooling water system operates, the lower are the make-
up water rate, blowdown rate and chemical consumption. This is highly desirable.
However, at higher COC, with its concomitant higher TDS in the cooling water, the
fouling (scaling) tendency of the cooling water becomes greater; this imposes a practical
maximum on the operating COC. For cooling water systems using Athabasca River water
in northern Alberta as fresh make-up water, for example, the operating COC is
maintained in the range of 3 to 5. Variation within this range is driven by the seasonal
change in the make-up water chemistry (especially dissolved calcium salts), the
effectiveness of the chemical treatment program for the cooling water and the knowhow
employed in the execution of the chemical treatment program. The chemicals used
consist of biocide, pH control agent, corrosion inhibitor, etc.

Limits on Cooling Water Supply (CWS) and Return (CWR) Temperatures

The returned cooling water is cooled in the cooling tower to CWS temperature. The CWS
temperature is a function of the ambient air temperature and humidity. In Alberta, the

104
CWS temperature is typically held at 24 °C, which corresponds to what can be optimally
achievable at the ambient air temperature and humidity “typical” on a summer day.

The CWR temperature, as seen by the cooling tower, is the combined temperature of the
cooling water outlets from all the water-cooled exchangers in the complex. As applied
locally to an individual exchanger, its cooling water outlet temperature is its own or local
CWR temperature. The limit on the local CWR temperature is typically set at 40 °C for
locations using Athabasca River water as make-up and for cooling water systems
operating at COC in the range of 3 to 5. This limit of CWR is considered optimal for
controlling the fouling tendency of the cooling water. Therefore, the cooling tower, the
supply and return headers and laterals, circulating pumps, and all ancillary (chemical
treatment facilities for the cooling water, etc) of the system are sized on the basis of these
CWS and CWR temperatures and of the combined heat duty of all water-cooled
exchangers in the complex. In a large complex such as a heavy oil upgrading complex or
a refinery, thousands of tonnes of cooling water per hour are kept in circulation to meet
the need of the cooling water users. The equipment and piping involved are enormous:
multiple pumps of hundreds of kW each and supply and return headers of large diameters
(in the order 48 inch to 72 inch) that run for kilometers on formidable support structures.
The cooling water system is therefore conceived and built to accommodate all planned
future expansion of the complex. The consequence of exceeding the capacity due to the
abusive use of CWR temperature can be very grave, as discussed below under the caption
The Game Played with CWR Temperature. Retrofitting the large complex with another
set of supply and return headers, circulating pumps and cooling towers to cover this
unforeseen shortage caused by the unchecked abuse of CWR temperature is hardly a
viable option. The only viable option is to nip any abuse, intentional or not, in the bud.

The Game Played with CWR Temperature

In process design the abuses of the CWR temperature are prevalent. Some of these abuses
are committed inadvertently due to the designer’s lack of appreciation of the
consequences of his action, while others are committed intentionally. Intentional abuses
are rampant in situations where the designer’s firm is contracted on a lump sum basis by
his client to design and to procure some pieces of equipment or entire operating units.

How does the use of a lower CWR temperature benefit the lump sum contractor? The use
of a lower CWR temperature (a) reduces of size of the water-cooled exchangers; (b)
enables the designer to use a single shell when two or more shells in series would
otherwise be required due to the need to circumvent a temperature pinch; (c) reduces the
size of other equipment associated with the coolers. For example, if the coolers are used
as interstage coolers of a compressor, a lower CWR temperature means the interstage
compressed gas streams can be cooled to a lower temperature thus reducing the size and
power of the compressor. In one episode, for the condenser of a refrigeration unit, the
designer of a lump sum contract used the CWR temperature of 30 °C instead the 40 °C
specified by the client. Given the CWS temperature of 24 °C, the temperature pick-ups
are 6 °C and 16 °C, corresponding to the CWR temperature of 30 °C and 40 °C,

105
respectively. It is easily seen that, for a given duty, the cooling water rate at 30 °C CWR
is 2.7 times (16 °C/6 °C) that at 40 °C CWR. The use 30 °C CWR enabled the refrigerant
to condense at a lower temperature than that which could be achieved with a CWR
temperature of 40 °C. This resulted in lower compressor power and a smaller refrigerant
compressor. In this same episode, the lump sum contractor also used a CWR temperature
of 30 °C (instead of 40 °C) for the inter-stage coolers of a large four-stage compressor.
This too resulted in lower compression power and a smaller compressor.

To make matter worse, the designer used CWR temperatures higher than 40 °C for other
water-cooled exchangers in the unit covered in the same lump sum contract, as high as 80
°C in some instances. In these other water-cooled exchangers, the hot process fluids only
need be cooled to temperatures well above 40 °C, such that the use of a high CWR
temperature does not significantly increase the size of the of the exchangers in question.
Of course the designer set the CWR temperature as higher as possible, but prudently not
so high as to compel him to use two shells in series to circumvent a temperature pinch.
Or, in the design parlance, the use of a higher CWR temperature only marginally reduces
the LMTD, but CWR used is not high enough to cause a temperature pinch and a bump
from a single shell to two shells in series. The purpose of using higher CWR temperatures
in the other exchangers (where sizes and cost are only marginally affected) is to
compensate for the extremely low CWR temperatures used in the major services (big
coolers associated with big machines). The compensation is such that the combined CWR
temperature from the entire unit comes closer to 40 °C (35 °C in this particular episode);
so the “optic” is not that bad in case anyone notices this game played with the CWR
temperature. It would require a rather strenuous stretch of the imagination to believe that
this elaborated scheme of manipulating the CWR temperature was conceived in the
client’s best interest. Unfortunately, this ruse of ‘robbing the proverbial Peter to pay Paul’
turns out to be successful more often than not, especially if the combined CWR
temperature is made close 40 °C, i.e. “almost” on the client’s target! The contractor bets
on the high probability that the client engineer wouldn’t pay too much attention to a
humble utility like cooling water, which appears to come from a limitless supply. If
allowed to proceed, the client will be left holding the proverbial bag in more ways than
one. More often than not, the lump sum contractor wins his bet!

This game is severely detrimental to the client in the following ways:

• At 35 °C CWR, for example, the temperature pick-up is 11 °C, based on the CWS
temperature of 24 °C. The cooling water requirement then turns out to be 1.5
times (16 °C/11 °C =1.5) that required if the client’s specification of 40 °C were
respected. (As already noted above at 30 °C CWR, the cooling water rate is 2.7
times that required at 40 °C CWR.) If designers of all units were allowed to
employ a ruse of this nature, it would not be long before the cooling water system
runs out of capacity - capacity which, be it recalled, is based on a CWR
temperature of 40°C, and prebuilt for planned future expansions.

• There are a few considerations that set CWR temperature at 40 °C. The salient
one is to control fouling due to scaling. As mentioned above, the fouling (or

106
scaling) tendency caused by calcium salt deposits depends on the TDS in the
water and on the effectiveness of the chemical treatment program of the cooling
water. More strongly, it depends on the water temperature. In the water-cooled
exchanger, one has to keep a tube skin temperature (on the cooling water side)
below a certain limit. This metal skin temperature is of course higher than the
bulk temperature of the water, typically 10 to 30 °C higher, depending on the heat
flux across the tube wall , and on the heat transfer coefficients and fouling
allowance on both sides of the tube. The conventional limit on this tube skin
temperature is 60 °C; this leads to the choice of 40 °C as the bulk CWR
temperature. This conventional limit of 60 °C almost a universal and it is
respected by scrupulous designers. There is some prudence implicit in this limit of
60 °C to allow for uncertainty in water chemistry. Calcium salt deposition occurs
precipitously and abruptly when the water temperature reaches a threshold. The
deposits adhere to the tube wall as scale. Because of the abrupt nature of the
occurrence, the loss of the cooling capacity of the affected exchanger can be
rather spectacular. One loathes to think what the tube metal skin temperature is
when the cooler is run at a cooling water bulk temperature of 80 °C; and how long
one can run the plant before it becomes inoperable due to the loss of capacity of
the fouled exchangers.

• By the use of a CWR temperature of 30 °C, condensation of the refrigerant can be


made to occur at a temperature close to 30 °C. This lower condensation
temperature means lower condenser pressure. As the refrigerant compressor
discharges against this condenser pressure, the lower pressure means smaller size
and lower power for the refrigerant compressor. This refrigeration unit will not
perform to the intended capacity when a CWR temperature of 40 °C is enforced in
operation.

Likewise, the 4-stage major compressor mentioned above will have its throughput
curtailed when the target CWR temperature of 40 °C is enforced for its inter-
coolers.

The upshot of all this is that the client is handed under-sized plants and under-
sized equipment, which is sure to have negative impact on plant throughput and
revenue.

107
Chapter 6:

THERMOSIPHON REBOILERS

6.1 Introduction

The thermosiphon reboiler is one in which the fluid flowing through it is driven by a
convective current, known sometimes as the thermosiphon effect. This convective current
arises from the density difference between the reboiler inlet and the reboiler outlet. The
reboiler inlet is the clear (or vapour-free) liquid from the last tray (namely, bottom-most
tray) of a fractionating tower plus internally recirculating liquid if any, and the reboiler
outlet is a mixture of liquid and vapour generated by boiling, hence the density
difference. The flow through the reboiler is maintained by an adequate liquid head
maintained in the inlet, known as the drive head.

In this chapter we present the major design considerations for the shell and tube
exchanger intended to be a thermosiphon reboiler. We will discuss the misconceptions
and uncertainties that some designers have about the requirements and the workings of a
thermosiphon reboiler.

For the purpose of this chapter, the fractionation tower is a device equipped internally
with trays or packings to facilitate liquid-vapour contact.

6.2 Types of Reboilers

Shell and tube reboilers used in fractionation come in three types: kettle, horizontal
thermosiphon and vertical thermosiphon. (As we are concerned here only with reboilers
of the shell and tube construction, we exclude from our discussion fired heaters used as
reboilers.) In terms of capital (equipment and installed cost), for the same reboiler service
and duty, the kettle is the highest and the vertical thermosiphon the lowest, with the
horizontal thermosiphon being intermediate. For the same service and duty, the tube
bundle size (tube area) is practically the same for all three types, any apparent difference
being within the margin of errors of the sizing calculations. The difference in capital cost,
therefore, arises from the complexity and amount of piping, shell size, foot print and
support superstructure for the reboiler. This can be inferred from Figure 6.1 (a), (b) and
(c).

The major disadvantage of a kettle reboiler is that it has to be placed at a considerable


elevation in order to gravity-drain the liquid from its shell back to the bottom of the tower
to avail of the liquid surge capacity provided there. The liquid trap-out compartment in
the kettle reboiler has too small a volume for the purpose of surge capacity. This
elevation requires a substantial superstructure. It is common to have to elevate a kettle
reboiler to 10 m above grade. Because of this elevation, the head room (vertical distance)
between the last tray and the tower bottom has to be increased, so that liquid from this

108
last tray can feed the elevated kettle reboiler by gravity. This extra headroom requirement
makes for a taller tower structure.

Liquid 1 Last tray liquid


trap-out 2 Vapour return
1 3 Hot fluid inlet
2 3 4 Hot fluid outlet
5 Liquid return

Bottom 4
surge E

GRADE GRADE

(a): Kettle reboiler. Elevation E usually requires the reboiler to have


its support superstructure. The large E is due to the need to
gravity-drain the liquid return back to the tower; and the the large
vapour return line requires large elbow room for its bends.
In some applications, E measures more than 10 m.

1 Last tray liquid 1 Last tray liquid


2 Tw o phase return 1 2 Tw o phase return
1
V 3 Hot fluid inlet 3 Hot fluid inlet
2 4 Hot fluid outlet V 4 Hot fluid outlet
3 L
L
2

Bottom Bottom 3
surge surge
4
E
E GRADE GRADE
GRADE GRADE 4

(c): Horizontal thermosiphon reboiler. The


(b): Vertical thermosiphon reboiler. The
horizontal shell rests on its saddles at
reboiler rest on its skirt or other
grade.
support at grade. It requires the least
foot print.
C ti li ith th t
Figure 6.1:
Reboilers: kettle, vertical thermosiphon and horizontal thermosiphon

For a thermosiphon reboiler, there is no elevation requirement for its placement; the
reboiler outlet (liquid plus vapour) is driven back to the tower by the thermosiphon effect
engendered in the boiling process. The outlet line (i.e. return line to the tower) is smaller
compared to the vapour return line of the kettle reboiler (for the same service and duty); it
requires less elbow room for its bends, hence a smaller foot print. The smaller outlet line
of a thermosiphon arises from the need to maintain an adequate vapour velocity in the

109
line to lift the returned liquid back into the tower. In the case of a vertical thermosiphon,
the return line can be a short straight horizontal run of piping with no bends, as suggested
in Figure 6.1 (b). As a consequence of this simple outlet piping the vertical thermosiphon
shell, though taller than the horizontal, does not always make for a taller tower than does
the horizontal, because its tallness is offset by the elbow room required for the more
complicated outlet piping of the horizontal thermosiphon. Note that in a vertical
thermosiphon reboiler the boiling liquid is placed in the tube side, because it is much
easier to achieve the desirable annular flow regime for the boiling liquid inside the
vertical tubes than it is in the shell side. In a horizontal thermosiphon reboiler the boiling
liquid is put in the shell side to accommodate the natural upward rise of the boiling
liquid.

6.3 Vaporization Ratio (or % Vaporization)

The vaporization ratio for a reboiler is the ratio (usually expressed as %) of the mass flow
rate of the vapour generated in reboiler to the total liquid mass flow rate entering the
reboiler. For a reboiler to function properly, there is a limit on % vaporization, beyond
which boiling becomes impaired. In the case of boiling on the shell side (kettle and
horizontal thermosiphon) this boiling impairment is due to vapor shrouding of the tube
surface, a flow regime that prevails at high vaporization ratios. Further generation of
vapour is hampered because, due to vapour shrouding, the boiling liquid is denied direct
access to the heat transfer surface. In the case of tube side boiling (vertical
thermosiphon), the two-phase flow regime inside the tube flips from annular flow to mist
flow when the vaporization ratio exceeds a certain threshold. In annular flow, an annular
layer of flowing liquid is held against the inside tube wall by the flowing vapour in the
core of the tube. This flow regime is favorable to boiling because the boiling liquid is at
all times in direct contact with the tube surface and the vapour generated is released to the
vapour flow in the core. In mist flow, however, the vapour phase is continuous with
dispersed droplets of liquid entrained in it. The entrained liquid droplets, which need to
boil, are not in direct contact with the tube surface. Further heat transfer from the tube
surface is curtailed and boiling stalls.

Most designers are not aware of such thing as a limit on vaporization as applied to
reboilers. Some designers, who are aware of such a limit, adopt a limit on vaporization of
30 %; but it is advisable to use 25 % as the prudent limit for both the shell side and tube
side boiling, because of the uncertainty incurred in the predication of flow regime
change. (See Sloley (1).) Between 25 and 30 % vaporization there is practically no
difference in tube area requirement; the drive head is only marginally higher at 25 %
vaporization (by about 350 mm liquid column) than it would be at 30 %. (Recall that the
drive head is the liquid head maintained over the inlet to the reboiler to sustain circulation
through the reboiler.) For a grass-root design this difference in the drive head is
insignificant. At 25 % vaporization we can be certain to stay clear of unfavorable flow
regimes that can severely curtail heat transfer and cause the reboiler to stall, resulting in
failure to meet fractionation targets. The apparent limit in a kettle reboiler should be
higher than 25 %. (See discussion in Section 6.4 below.)

110
For example, if a reboiler is mistakenly designed with 70 % vaporization, it will not be
able to generate enough vapour to meet this ratio. (We say mistakenly designed, because
the designer is not aware of the limit on vaporization.) Further boiling stalls when the
vaporization ratio reaches 35 %, for example, because of vapour shrouding (shell side
boiling) or the prevailing mist flow regime (tube side boiling). Consequently vapour is
generated at only half the intended rate. The impact on fractionation of this shortfall in
vapour generation will be severe. It should be noted that most heat exchanger and
fractionation simulation and sizing programs do not take into account the adverse effects
of vapour shrouding and mist flow on heat transfer; they do not recognize the limit on
vaporization, let alone warn the user if this limit is exceeded.

6.4 The (Wrong) Reasons Why A Kettle Reboiler Is Selected

It is curious why the kettle reboiler is used at all given that it has no operating advantage
to offset its higher capital cost and its inordinate encumbrance on plot area. However, it is
often conjectured by its proponent that the kettle reboiler can tolerate a higher apparent
vaporization ratio than the thermosiphon.

A Vapour
return Kettle
Hot fuid shell
Convective
Tube current
bundle

Hot fluid Inlet Liquid Inlet


(last tray return
A
liquid) View A-A

Figure 6.2:
Flow through the tube bundle induced by convection reduces the apparent %
vaporization by increasing the liquid flow "seen" by the tube surface, which equals
convective current plus liquid inlet to the reboiler. See View A-A.
Note, for a given tube bundle, the inordinately large shell characteristic of the kettle
reboiler due to the need for vapour-liquid separation.

For the purpose of the discussion in this section, we define the apparent vaporization
ratio as the mass flow rate of vapour leaving the reboiler to the mass flow rate of liquid
feed to the reboiler; and we define local vaporization ratio as the mass flow rate of
vapour generated to the local liquid flow rate across the tube bundle; this local liquid
flow rate is the sum of the liquid feed to the reboiler plus the convective current of
recycled liquid induced in the reboiler. By definition, therefore, the local vaporization
ratio is always smaller than the apparent vaporization. The kettle reboiler consists of a
tube bundle immersed in a pool of liquid. A natural convective circulation of the liquid in
the pool is induced by boiling. This convective circulation is additive to the inlet liquid
flow to the reboiler. The total liquid flow rate across the tube bundle is thereby increased.

111
(See Figure 6.2.) Clearly it is the local vaporization ratio, i.e. the actual vaporization ratio
”seen” by the tube surface, that is operative in affecting boiling. Therefore, if a limit of
25 % is imposed on this local vaporization ratio, then the corresponding apparent
vaporization must be greater than 25 %. However, it is uncertain to what extent this
convective recirculation contributes to the liquid flow across the bundle, and how much
above the 25 % can be allowed for the apparent vaporization ratio as a design parameter
for the reboiler. Hence a designer that exceeds 25% apparent vaporization ratio in a kettle
reboiler is entering unknown and therefore risky territory. A 25% apparent vaporization
rate remains a prudent design limit for a kettle reboiler, assuming that one ever has to
design a kettle reboiler!

The following may be the reasons why some designers stick to the kettle reboiler
tenaciously to the exclusion of the thermosiphon reboiler:

• In the printout of a fractionation simulation software the reboiler function is


usually represented by an icon or symbol that bears the likeness of the kettle
reboiler. This representation is strictly symbolic and we surmise that the software
developer has no intention of dictating the use of the kettle reboiler as the only
physical device for generating the vapour required. Also in the same printout, the
liquid feed from the fractionation tower to the reboiler is shown, symbolically, as
being fed exclusively from the bottom-most fractionating tray directly into the
reboiler. (See Figure 6.1 (a).)Neither the type of reboiler nor how it is fed from
the tower has any bearing on the fractionation simulated by the software. As far as
the design of the physical reboiler is concerned, and how it is physically fed from
and connected to the tower, what matters in the fractionation simulation results
are the absolute vapour generation rate required of the reboiler and the
accompanying heat duty of the reboiler. It is unfortunate that this symbolic
representation of the reboil requirement (i.e. vapour generation rate and reboiler
duty) is taken literally and inspires many designers to use the kettle reboiler
exclusively and connect it to the tower as shown in Figure 6.1 (a).

• Some designers believe that the thermosiphon design incurs a technical risk
because of the “delicate” hydraulic balance that it requires for its functioning.
Therefore it is something to be avoided at all cost. This fear of the thermosiphon
reboiler is hardly justifiable. In fact, the operation of the thermosiphon reboiler is
inherently stable, in spite of the seemingly delicate hydraulic balance involved.
This point will be elaborated in Section 6.8. Contrary to popular belief, the
thermosiphon is actually inherently self-correcting: it recovers from perturbations
on its own, without any external countervailing control action.

• The choice of the kettle reboiler is made on the perceived higher tolerance of
vaporization ratio, as discussed above. (We are being generous in citing this as a
motivation for the designer to select the kettle reboiler, because most advocates of
the kettle reboiler are not cognizant of the existence of the limit on vaporization.)
Be that as it may, this hardly justifies the higher cost and encumbrance of the
kettle reboiler because the limit on vaporization ratio can be easily circumvented

112
by using a recirculating arrangement with the thermosiphon reboiler, as will be
seen below in Section 6.5.

• Some design organizations use the kettle reboiler exclusively because of their
long entrenched practices. They exploit their client’s fear of a “technology step-
out” and of “interfering with the contractor’s internal practice, work processes and
in-house expertise”. This duet between the design organization and client quite
often ends up ruling out a less expensive thermosiphon reboiler.

As there is no further justification for the use of the kettle reboiler, the discussion in the
remaining of this chapter will be focused on the thermosiphon reboiler.

6.5 Once-through versus Recirculating Thermosiphon Reboilers

The once-through arrangement is one in which the liquid feed to the reboiler comes
solely from the bottom-most fractionating tray; and there is no other liquid stream being
recirculated to the reboiler. This simple arrangement should be used whenever feasible,
i.e. within the limit of the vaporization ratio discussed above.

In the example in Figure 6.3, the results of a fractionation simulation (of a sour water
stripper) show that the liquid rate from the last tray is 207,071 kg/h; and the reboiler is
required to generate 35,202 kg/h of vapour to satisfy the fractionation requirement, with
the accompanying reboiler duty of 69.7 GJ/h. We see that a once-through arrangement
for feeding the reboiler with liquid is indeed feasible, as this entails a vaporization ratio
of 17.0 %, comfortably below the limit of 25 %. This simple once-through arrangement
should be therefore adopted.

L1 207,071 kg/h

V0
V 0 35,202 kg/h
L0 L0 171,869 kg/h

Therm osiphon
Reboiler
Duty QR
Bottoms 69.7 GJ/h
171,869 kg/h

V0 35,202
Vaporization = = = 17.0% < 25 %
L1 207,071
Figure 6.3:
Feasible once-through arrangement: a once-through arrangement
is feasible here. The vaporization as "seen" by the reboiler is 17.0
% (< the limit of 25 %)

113
However, in another example (a sponge oil absorber for a gas recovery unit) as shown in
Figure 6.4, the fractionation requirements (last tray liquid of 529,370 kg/h and vapour
generation of 221,100 kg/h) clearly cannot be met by a once-though arrangement, as this
would entail a vaporization of 41.8 %. A downcomer and a baffle are used to induce a
liquid recirculation of 355,030 kg/h, so that the vaporization ratio (as “seen” by the
reboiler) is brought down to 25 %.

L1 529,370 kg/h L1 529,370 kg/h


V0
V0
V 0 221,100 kg/h V 0 221,100 kg/h
L0 L0
L0 308,270 kg/h
L0 663,300 kg/h
Therm osiphon LR
Therm osiphon
Reboiler
Duty QR Reboiler
Bottoms Duty QR
102.3 GJ/h
308,270 kg/h 102.3 GJ/h
Bottoms LR + L1
308,270 kg/h 884,400 kg/h
V0 221,100
Vaporization = = = 41.8% > 25 % For LR = 355,030 kg/h :
L1 529,370
Vo 221,100
Vaporization = = = 25.0 %
L R + L1 355,030 + 529,370

(b) Preferential recirculating arrangement:


(a) Infeasible once-through arrangement:
Inducing liquid recirculation LR to achieve 25 %
This arrangement would result in a
vaporization ratio of 41.8 %. vaporization

Figure 6.4:
Recirculating liquid to reduce vaporization to 25 %

For the recirculation arrangement, two variations are used: recirculation with baffle and
recirculation without baffle. (See Figure 6.5 (a) and (b).) When the tower has a baffle a
constant liquid head (drive head) is maintained, since liquid is always overflowing the
baffle into the inventory of bottom product at the bottom of the tower. The drive head is
thus de-coupled from the level in the bottom product inventory, which is free to vary
without affecting the drive head. The baffle height is designed to drive the right amount
of liquid back into the reboiler, taking into account the total head loss in the circuit due to
friction, fittings, entry, exit, etc. This flow sustains the two phase flow from the reboiler
back to the tower. The use of a baffle is by far the better design. Without a baffle, on the
other hand, the drive head would have to be provided by the bottom product level. This
means that the inventory level in the bottom of the tower must be held constant for the
sake of the drive head; this is undesirable because it completely defeats the purpose of the
bottom product compartment, which is to provide surge capacity to the bottom product

114
for control and safe operation. A variation of level in the bottom product compartment
affects the hydraulics, and consequently the functioning of the reboiler.

There is another but seldom appreciated advantage of the recirculation with baffle
arrangement as shown in Figure 6.5 (a). This arrangement is also known as preferential
recirculating, so called because the entire liquid flow from the last (bottom-most)
fractionating tray is preferentially routed to the reboiler. This allows the reboiler’s
function to approach as nearly as possible a theoretical tray; whereas in the recirculation
without baffle arrangement, the last tray liquid is completely back-mixed with the bottom
product. This is detrimental because back-mixing reduces the mass transfer potential and
negates the approach of the reboiler to a theoretical tray. Also, as can be appreciated in
the case of recirculation without baffle, a portion of the last tray liquid is mixed and
withdrawn with the bottom product, thus skipping the reboiler completely.

Last tray
liquid Liquid in this compartment One single level for
finds its ow n level; product surge and
Reboiler Last tray
any bubbling action drive head for reboiler.
return liquid
here has no effect (A highly undesirable
(2-phase) Reboiler
on the constant situation.)
return
Breakout apron clear liquid drive.
(2 phase)
Constant drive Backmixing of last Returned
Overflow Dow ncomer extended
head for reboiler tray liquid and vapour
to effectively segregate Returned
returned liquid in
last tray liquid from liquid
Drive bottom product and in
Bottoms baffle returned liquid feed to reboiler
Reboiler
Bottom surge level Recirculated liquid
Bottoms feed
decoupled from (clear liquid)
drive head Reboiler feed

(a) Recirculation with baffle; also known as (b) Recirculation without baffle
preferential recirculation

Figure 6.5:
Recirculation arrangement: with and without baffle

Contrary to a mistakenly held notion, the choice between the once-through and
recirculating arrangement is not a matter of the designer’s style or preference; nor can it
be mandated as standard design practice of any organization. It is interesting to note that
some organizations adopt the once-through arrangement exclusively as their design
practice, sometimes touted as their forte, without regard to the limit on vaporization ratio.
One may surmise that this design practice may have been inspired by the fact the printout
of a fractionation software always shows a once-through arrangement, regardless of
vaporization ratio. As pointed out earlier, the reboiler and its connections to the tower as
indicated in such a printout are meant to be symbolic, not to be taken as prescriptive of
the actual physical device and set-up to meet the reboil requirement.

115
The choice between once-through and recirculating arrangements is dictated solely by the
limit on vaporization ratio. The purpose of the recirculating arrangement is to reduce the
vaporization ratio as “seen” in the reboiler, whenever it is necessary, to keep this ratio
within the limit of 25 %. See example in Figure 6.4 above. Personal preference and style
and company policy have nothing to do with this choice.

6.6 Simulation of a Preferential Recirculating Thermosiphon with Fractionation

When a reboiler is designed for recirculation, the composition of the vapour, at the same
rate as dictated by fractionation, differs slightly from that of a once-through arrangement
based on which the fractionation has first been simulated. Once it is realized that a
recirculating arrangement is required, the fractionation simulation can be repeated with
the recirculation arrangement incorporated in the scheme, in order to yield more rigorous
data (thermodynamic and transport properties) for the sizing and design of the physical
reboiler. It should be noted that in this repeated simulation, the purpose of which is to
represent the preferential recirculating arrangement rigorously, all fractionation targets
are met exactly as before when the simulation was based on once-through reboiling.
Figure 6.6 shows the set-up for this rigorous simulation. Some fractionation programs
offer a built-in facility to simulate fractionation with a recirculating arrangement for the
reboiler. The user should carefully examine this built-in facility to see if it corresponds to
the correct scheme shown in the Figure 6.6.

Colum n
Extraneous operations:
operation
Heater
Feed Separator2 Recycle
Tee Mix
Last tray liquid

Feed 1 Vapour return


(Vapor return)
Separator2
Heater
(Reboiler) True bottoms
Colum n
Mix Recycle Tee
Figure 6.6:
Simulation of preferential recirculating thermosiphon; this requires five unit operations
extraneous to Column in order to properly constitute the liquid feed to the reboiler.

The reboiler is Heater. Column is reboilerless, with two feeds. Vapour return
becomes a Column feed (Feed 1). Separator2 liquid is recycled to Heater and
drawn as true bottom product. The names (in boldface) of these unit operations are as
used in the commercial software Aspen HYSYS (AspenTech).

116
6.7 Vertical versus Horizontal Thermosiphon Reboiler

It was pointed above that the use of the kettle reboiler cannot be justified and its use is
misguided. Now the designer is left with the choice between the vertical and horizontal
thermosiphon. It should be emphasized here that this choice has nothing to do with the
once-through or recirculating arrangement, which arrangement applies to both the
vertical and horizontal thermosiphon and is dictated solely by the limit of 25 % imposed
on the vaporization ratio. The vertical thermosiphon requires less foot print and has
simpler piping than the horizontal. See Figure 6.1 (b) and (c). For a given service and
duty, the tube area is the practically the same for either vertical or horizontal. However,
the shell of the horizontal thermosiphon has to be larger because the tube bundle has to
have the NTIW (no tubes in window) geometry to provide header space above and below
it in the shell. See Section 5.5.3 on the NTIW tube bundle. The vertical thermosiphon
should always be used unless the following factors, which favour the horizontal
thermosiphon, become predominant:

• The heating medium has a high fouling tendency

The heating medium may be a process fluid with high fouling tendency; and it
is preferable to have the more fouling fluid on the tube side because cleaning
the inside of the tube is easier than the outside during a shutdown. In a
horizontal thermosiphon, the heating medium flows in the tube side.

• An inordinately long tube bundle is required for the vertical thermosiphon

In a vertical thermosiphon the boiling liquid flows vertically up in the tube side
of a single-pass tube bundle. An optimized design of a vertical thermosiphon
may be one that results in an inordinately long (tall) tube bundle. For example,
the hot fluid, flowing in the shell side, may be a liquid with inherently poor heat
transfer properties (high viscosity, low thermal conductivity of the liquid). A
long (tall) shell would be required to produce an acceptable heat transfer
coefficient in the shell side. The fractionation tower skirt will have to be higher
in order to accommodate this tall bundle. The higher cost as a result of the tall
bundle may not be off-set by the advantages of the vertical thermosiphon, such
as a small footprint and simple piping connecting it to the tower.

It should be noted here that the length of the tube bundle has nothing to do with
the throughput of the reboiler. To accommodate higher throughput, one
increases the number of tubes (i.e. bundle diameter), not the tube length. The
tube length (height) of vertical thermosiphon is determined by the number of
tubes to be used to induce an appropriate vapour velocity in the tubes, so that
the annular flow regime can be sustained in the upper part of the tubes; with the
number of tubes so determined, the length (height) of the tubes is given by the
area required. Due to this requirement of the annular flow regime in the tubes,
higher reboiler throughput (requiring greater tube area) must be accommodated

117
by increasing the number of tubes (not their length) in order maintain the same
velocity that sustains this favourable flow regime

In the process industry, one comes across many instances where the choice between the
vertical and horizontal thermosiphon is not necessarily made on the above two criteria.
Rather, the choice is made for the following wrong reasons:

• The horizontal thermosiphon is the only one sanctioned in the design


organization’s design practice. This is supposedly based on the cumulative
experience of its technical staff, and its supposedly enormous wealth of data on
this particular design. It has well established methods and procedures; and much
of the work process associated with the well established practice is
“automated”. Doing anything different would be a step-out from the standard
and is usually frowned upon in an organization. Here again, the client’s fear of
“technology step-out” and “disrupting the contractor’s established work
process” is unabashedly exploited.

Pull bundle Pull bundle


channel head channel head
TEMA A or B TEMA C
Stationary Stationary
tube sheet tube sheet

tube Tube
shell
Shell
gasket
packing

Floating Floating
tube sheet tube sheet
channel head Channel head
TEMA P, A or TEMA P, A or
B modified B modified
Figure 6.7:
Removable single-pass tube bundle for vertical thermosiphon

• Some designers claim that the vertical single-pass tube bundle has to be on fixed
tube sheets, such as TEMA type NEN. (See any shell and tube exchanger
design handbooks or TEMA publications for the designations of TEMA type.)
This exchanger type is to be avoided because the fixed tube sheets have little
resilience against thermal expansion, and because the tube bundle cannot be
removed for maintenance and cleaning. This claim is patently false, but it is
made to sound so technical and so expertly articulated that, more often than not,
the client, discouraged from pursuing the matter further, simply acquiesces! The
claim is false because a single-pass tube bundle does not have to be of the fixed
tube sheet type. Also, the bundle can be made removable just as easily as the

118
bundle of the horizontal thermosiphon. See Figure 6.7. It is interesting to note
the length to which a designer would go, including resorting to disinformation,
just to stay within the comfort zone of the standard design of his organization.

• The duty is too high; so the use of the vertical thermosiphon would require a
prohibitively tall reboiler. As discussed above, higher duty is accommodated by
the use of more tubes, not longer (taller) tubes.

6.8 Inherent Stability of the Thermosiphon Reboiler

The process dynamics of the thermosiphon reboiler have the very desirable characteristic
of being self-stabilizing. This means its operation is inherently stable: when displaced by
perturbations from a steady state, it responds by always settling itself into another stable
steady state promptly without any external corrective control action; it does not run away.

Last tray Last tray


liquid liquid Chimney tray
Chimney tray

Operating
Reboiler return L1 L1 Operating
range of last
(2-phase) range of
tray liquid
Reboiler return last tray
Drive level
(2-phase) liquid level
head providing
(fixed) Dow ncomer
drive head

Bottoms Drive baffle L R Bottoms

Reboiler
Reboiler
Reboiler feed Reboiler feed
= L1 + LR = L1

(a) Preferential recirculating (b) Once-through

Figure 6.8:
Inherent stability of the thermosiphon reboiler

To see this inherent stability, let’s look at the preferential recirculating thermosiphon
depicted in Figure 6.8 (a). The drive head is fixed as given by the height of the drive
baffle. (Note that the crest height of the overflow over the drive baffle is insignificant in
terms of the hydraulics of the thermosiphon and can be ignored; by contrast, for example,
the crest height over the aqueous overflow weir in an oil skimmer is highly significant
and must be taken into account, as discussed in Chapter 4.) The last tray liquid L 1 from
the fractionation tower rides freely on this fixed head, due to the extended downcomer
from the draw-off tray; this last tray liquid finds its own level in the range afforded by the
downcomer space and the liquid draw-off tray (the chimney tray). This means that the

119
last tray liquid always assumes whatever level is necessary, so that the entire last tray is
routed preferentially to the reboiler. The head difference between the drive baffle and the
reboiler drives the recirculating liquid rate L R . If heat input to the reboiler increases, the
fluid density in the reboiler decreases due to increased vapour, and the flow rate L R
increases. Similarly, a lower heat rate will automatically lead to a decrease in L R .The
recirculating liquid rate L R will also adjusts automatically to fluctuation of the last tray
liquid rate L 1 .

The self-stabilizing dynamics of the once-through arrangement is shown in Figure 6.8


(b). The drive head is the level of the last tray liquid, which is free to vary within a range.
At a given reboiler heat input, the drive head settles to a level sufficient to drive the flow
through the reboiler. If the heat input increases, the return two-phase fluid from the
reboiler becomes less dense because of increased vapour in it, and the level of the drive
head drops automatically to compensate. If the reboiler heat input decreases, vaporization
is thereby reduced making the return fluid denser; then, in order to maintain flow through
the reboiler, the drive head level rises to countervail the higher head exerted by the
reboiler return due to its increased density. This drive head also automatically adjusts
itself to maintain flow through the reboiler in response to other perturbations, such as
fluctuations in the last tray liquid rate.

In spite of its inherent stability, the impression still persists among many designers that
the thermosiphon reboiler is “very delicate” and calls for specialized design and operating
skill. To find out how much drive head is needed, the static head sustained in the return
line plus head losses due to friction and fittings, etc, must be calculated. This means
every fitting, entry nozzle, exit nozzle, bend and tee in the inlet and outlet piping must be
taken into account. This attention to the details of piping and fittings may be what gives
the false impression that the working of the thermosiphon is too delicate or “dicey”. Well,
this has nothing to do with delicateness; it is a question of significance: every fitting is
significant because of the magnitude we work with, since the total head loss incurred in a
thermosiphon is in the order of 150 to 200 mm of drive head required. By contrast, for
example, in a long pipe which entails a large pressure drop (1000 kPa, say) due to the
sheer length of the flow path, the head loss incurred by a bend (or any fitting), or even by
a large number of bends and fittings, is then quite insignificant. We can afford to miss
quite a few of them in these hydraulic calculations without affecting the final result
significantly.

A system with delicate dynamics, on the contrary, is one which can become unstable or
erratic when it is perturbed, or one that has very small tolerance to perturbations. As we
indicated above, the thermosiphon reboiler has inherently stable dynamics; it is not
delicate but robust, even though a fitting or a bend happens to be significant in its
hydraulics. Significance is relative; it nothing to do with robustness or the inherent
stability of a system.

Because of the inherently self-stabilizing dynamics, we can always add a comfortable


margin to the drive head without fear of it over-driving the flow. For example, if the
calculation shows that we need a drive head of 2800 mm, we can add a margin of say 500

120
mm to make in 3300 mm in the final design. This margin is for us to err on the good side,
of course. The extra liquid recirculation through the reboiler due to this margin will
always be countervailed by a higher friction loss and a higher reboiler outlet head. The
higher friction loss is simply due to higher flow. The higher reboiler outlet head is due to
the fact that, with the extra liquid recirculation, the vapour ratio in the reboiler outlet fluid
decreases causing an increase in the fluid density, while the vaporization rate remains the
same. In short, the extra margin for the drive head cannot cause the recirculation through
the reboiler to run away.

6.9 Chimney Tray, Liquid Trap-out and Draw-off

In the sketches in this chapter, we have shown the use of the chimney tray to facilitate
liquid draw-off from the fractionation tower, which draw-off is routed in its entirety
through the reboiler; we have also on occasions shown liquid draw-off made from an
active fractionation tray, as in Figures 6.1 and 6.5 b. We note here specifically that liquid
draw-off can be made directly from a fractionation tray, as shown in Figure 6.9 (a) and
(b). The chimney tray does not perform any fractionation function since it provides no
contact of the vapour rising from below it with the liquid flowing on it. However, the use
of the chimney tray for liquid draw-off is an option that has some considerable
advantages over direct draw-off from a fractionating tray:

• As the tower is charged with liquid in preparation for start-up, the reboiler is
self-priming because the chimney tray traps out all liquid descending the tower
and routes it all through the reboiler; whereas during this charging process, a
fractionation tray cannot be relied on to prime the reboiler; this is so because,
during the charging process at start-up, there is no upflow of vapour through a
fractionating tray to sustain, reliably, a liquid level on the tray sufficiently to
route the liquid to the reboiler. For this reason a start-up line must be provided
to positively prime the reboiler with liquid for start-up, as shown in Figure 6.9.

Tray 2 Tray 2
liquid liquid
Tray 1* Tray 1* Liquid
Reboiler draw -off
return Vapour to reboiler
Vapour
Liquid
Reboiler Liquid
Start-up** return
Start-up**

Bottoms Bottoms

Liquid draw -off *Tray 1 is the last (bottom-most)


+ recirculation to fractionation tray
reboiler **Start-up line: to reboiler inlet for
priming the reboiler
(a) Liquid draw-off plus recirculating liquid (b) External liquid draw-off
Figure 6.9:
Liquid draw-off directly from an active fractionation tray
121
• During operation, a malfunctioning fractionating tray used as draw-off tray may
leak and load the reboiler with insufficient liquid flow, or not at all. The leaked
liquid is mixed with the tower bottom product and, therefore, bypasses the
reboiler. This bypassing is detrimental to fractionation. The chimney tray, on
the other hand, assures complete liquid trap-out and thus always loads the
reboiler with the last-tray liquid in its entirety (no leakage possible) (see Figure
6.8);

• A chimney affords better vapour disengagement. There is desirable especially


when it is used to draw off liquid to feed a pump, such as a side stream draw-off
from a fractionation tower.

It is recognized here that the advantages of the chimney tray far outweigh its cost; it
has been included in the fractionation towers with greater frequency in the past two
decades. It should be clear that the use of a chimney tray is not a prerequisite of a
thermosiphon reboiler, although its use is preferable to that of a fractionation tray
for liquid draw-off.

122
Chapter 7

THERMODYNAMICS – AN OVERVIEW

7.1 Introduction

This chapter is a short overview of the first and second laws of thermodynamics. In this
overview, these two laws will be “re-constructed” in a manner somewhat different from
what is usually presented in most text books. The aim here is to see how all those
expressions we use in our analyses and calculations come about and, hopefully, to show
them to be less abstruse than they are made out to be. We emphasize that there are only
two fundamental state functions in thermodynamics: internal energy U and entropy
S . Together they capture all thermodynamic relations in their entirety. By this we hope to
connect with the science of thermodynamics at the fundamental level and to dispel the
mystery enshrouded in a maze of derived functions and relations.

The thermodynamic principles underlying the refrigeration process will be discussed in


Chapter 8. Any other process is amenable to this thermodynamic treatment, but we
choose refrigeration because it is familiar and because it manifests these principles in all
transparency, simplicity and elegance. We could skip this overview and, right away,
apply the laws of thermodynamics, using the expressions of these laws in their highly
specialized and deeply truncated forms that cater to the refrigeration process. This
approach, however, is not helpful because it would be just another repetition of the
presentations already given in other text books. We would then simply relate
thermodynamic quantities appearing in the refrigeration process to the specialized and
truncated expressions of the laws. This is tantamount to stating the results without saying
how we arrive at them.

7.2 First law

The first law of thermodynamics is the precise statement of the law of conservation of
energy. Before we formulate it, we need to state the premises:

System

The system comprises of a collection of processes within its boundary.

Surroundings

The surroundings consist of everything outside the system boundary - everything that
interacts directly with the system through the exchanges of material substances, heat and
work.

123
A succinct high-level statement of the first law would be: the energy transactions brought
about by the exchanges of material, heat and work between the system and its
surroundings always balance; no energy is created nor destroyed in the transactions.

Let’s examine the energy transactions with an example. For its familiarity, we consider a
system consisting of a steam turbine and its lube oil ancillary as shown in Figure 7.1.

Surroundings

System 2 **
Material m 2

m 1 Turbine Shaft w ork w S 2


Material 1 **
Heat q 2
Heat q1

Shaft w ork w S 1 Lube Pump


Ancillary

** For w ork transactions at 1 and 2 ,


see Figures F.2 and F.3.
Figure 7.1:
First law of thermodynamics; or statement of the conservation of energy

Here the subscript 1 denotes material, heat and work entering the system (from the
surroundings) and the subscript 2 denotes material, heat and work leaving the system.
The subscript c denotes material and energy held inside the system. The properties,
conditions and flow rates of the material (steam, in this example) at the point 1 (entry)
and point 2 (exit) and inside the system are:

Point 1: Point 2 Inside system


Mass flow rate, kg/s: m 1 m 2
Internal energy** of material, J/kg: U 1 U2 UC
3
Specific volume of material, m /kg V1 V2
Pressure, Pa (abs): P1 P2
Velocity of material, m/s: u1 u2
Elevation, m: z1 z2
Hold-up of material, kg: mC

** A material substance has internal energy U by virtue of its constitution and by virtue
of the motion of its constituents; an elaboration of this is made in Section 7.7.

We are ready to tally up the energy. But before doing so, let’s look at the work entering
and leaving the system due to the flowing material, other than the shaft work w S 1 and
w S 2 indicated in Figure 7.1.

124
Work other than shaft work

Work is done to the system by the surroundings at the rate of m 1 P1V1 (at point 1) due to
material flowing in or pushing its way into the system; and work is done by the system to
the surroundings at the rate of m 2 P2V2 (at point 2) due to material flowing out or pushing
its way out into the surroundings. These work quantities are in addition to the shaft
work w S 1 , w S 2 indicated in Figure 7.1. They are shown separately in Figures 7.2 and 7.3
for the sake of clarity.

Volume of material entering in time interval δ t : δ m1V1


Surroundings System
Work done TO system due to material δ m1 P1V1
1 flow in time interval δ t :

 δ m1  dm1
Instantaneous w ork done TO  P1V1  = P1V1
Mass δ m1 of material
system due to material flow :  δt  δ t →0 dt
entering in time interval δt
1 1 1 = m P V
Figure 7.2:
Work done TO system due to material flowing in (or material pushing its way into the system)

Volume of material leaving in time interval δ t : δ m2V2


System Surroundings
Work done BY system due to material
flow in time interval δ t :
δ m2 P2V2
2

 δ m2  dm2
Instantaneous w ork done BY
 P2V2  = P2V2
Mass δ m2 of material
leaving in time interval δt
system due to material flow :
 δt  δ t →0 dt
= m 2 P2V2
Figure 7.3:
Work done BY system due to material flowing out (or material pushing its way out against the surroundings)

Energy tally

The tally of the energy terms entering the leaving the system is straight forward:

Entering Leaving
[All entries are in J/s (i.e. watt)]
Internal energy of material: m 1U 1 m 2U 2
2
u u 22
Kinetic energy of material due to bulk flow: m 1 1
m 2
2 2
Gravitational potential energy of material: m 1 z1 g m 2 z 2 g

125
Heat: q1 q 2

Work: w S 1 + m 1 P1V1 w S 2 + m 2 P2V2


d
Energy accumulation in system: (mCU C )
dt

Formulation of the First Law of Thermodynamics

The first law (energy conservation) simply requires:

Energy entering system = energy leaving system + energy accumulation inside system.

From our tally we can formulate the first law as:

 u12 
m1 U 1 + P1V1 +
 + z1 g  + q1 + w S 1 =
 2 
Energy entering system

 u 22  d
m2 U 2 + P2V2 +
 + z 2 g  + q 2 + w S 2 + (mCU C )
 2  dt
Energy
Energy leaving system
accumulation
inside system

Rearranging:

 u 22   u12  d
m2 U 2 + P2V2 +
 + z 2 g  − m1 (U 1 + P1V 1 ) +
 + z1 g  = (q1 − q 2 ) − (w S 2 − w S 1 ) − (mC U C )
 2   2  dt

Or:

 u2   u2  d
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g  = q − w S − (mC U C ) (7.1)
 2   2  dt

In the above expressions, we considered only one stream each of material, heat and shaft
work entering and leaving the system, for simplicity. The tally can be extended to any
number of such streams. Also, we have used an arbitrary system consisting of familiar
processes: steam turbine and its ancillary, for easy appreciation. This arbitrary system in
no way restricts the generality and validity of (7.1). For example, if we use some process

126
which entails material transformations (chemical reactions), the expression (7.1) remains
valid. Instead of a pure material stream (steam) entering and leaving, we will have a
mixture of reactants entering and a mixture of reaction products. The internal energy U
and specific volume V will then pertain to the mixture, instead of a pure component
(steam). The same is said of the material hold-up, where m C and U C pertain to those of a
mixture. In Sections 7.6 and 7.7, the internal energy U of a mixture and the ∆U
associated chemical reactions are introduced.

Sign convention for heat and work

We have used the substitution: q = q1 − q 2 and w S = w S 2 − w S 1 . From this it can be seen
that q is the net heat entering the system; while w S is the net work leaving the system.
Implicit in this substitution is the sign convention that:

• The numerical value of heat entering the system is positive; while that of heat
leaving the system is negative; and
• The numerical value of work leaving the system is positive; while that of work
entering the system is negative.

This is the convention adopted in most process engineering texts. Note that this
convention is by no means universal. A different sign convention is frequently
encountered in texts used in other disciplines of science and engineering.

7.3 Specialization of the First Law

The general expression (7.1) can be specialized to specific situations. These specific
situations fall into two broad categories: (1) a steady state flow process and (2) a closed
process. It should be noted the specialized relations so obtained remain valid, regardless
of which category a situation may fall into. In other words, a relation obtained from
applying the first law to a closed process will also be valid for an open steady state flow
process, and vice versa. Which relation we choose to use in a particular situation is
strictly a matter of convenience, not of validity. This point will be elaborated below when
we have derived these specialized expressions.

7.3.1 Steady State Flow Process

Steady state means that the state functions, such as U and S, and material properties at
any point in the system do not vary with time – although they may vary from one point to
another. Also, in a steady state flow process, there is no accumulation of material and
energy inside the system. This requires:

127
d
(mCU C ) = mC dU C + U C dmC = 0 (No accumulation of material and energy:
dt dt dt
dmC dU C
= 0; = 0)
dt dt

m 1 = m 2 = m (Requirement of material balance)

The first law equation is then reduced (but not truncated, yet) to:

  u2 u2  
m (U 2 + P2V2 ) − (U 1 + P1V 1 ) +  2 − 1  + ( z 2 − z1 )g  = q − w S
  2 2  

Or:

 u2 
m ∆(U + PV ) + ∆ + g∆z  = q − w S (7.2)
 2 

Here, in the interest of brevity, we have introduced the difference operator ∆ to mean that,
for a given quantity A , ∆kA = k∆A = k ( A2 − A1 ) , where k is a constant. In other words,
∆A expresses the difference between the value of A at point 2 (or state 2) and its value at
point 1 (or state 1).

Furthermore, we write this already condensed relation (7.2) by expressing it on the basis
of per unit mass of the material transacted and by using the symbol H for the
cluster U + PV :

u2 q w
∆(U + PV ) + ∆ + g∆z = − S
2 m m

Or:

u2
∆H + ∆ + g∆z = Q − WS (7.2a)
2

Where: Q = q m and Ws = w S m are the net heat entering and net work leaving the
system per unit mass of material transacted; H (called enthalpy) serves as shorthand for
the cluster U + PV .

Truncation

We have finally arrived at the equation for the first law applied to the steady state flow
process (7.2) and its shorthand version (7.2a). We have done so by using substitution,

128
shorthand and generalization to the basis of unit mass of material transacted. Note that so
far we have not resorted to approximation by deleting any terms that we may consider
insignificant. In other words, (7.2) or (7.2a) above, though condensed, is nonetheless
exact and complete for the steady state flow process.

In many processes, the terms involving the kinetic energy of bulk motion and
gravitational potential energies, ∆ u 2 2 and g∆z , respectively, are insignificant. These
terms can be dropped to give us the truncated or approximate equation:

∆H = Q − WS (7.3)

Now we finally have this equation (7.3), the ubiquitous work-horse equation of process
design and analysis. We must, however, remain mindful of how we have landed on it.
This work-horse equation has been specialized in two steps: reduction followed by
truncation of the terms deemed insignificant: ∆ u 2 2 g and g∆z . The reduction step,
whereby the terms whose values are zero are eliminated, gives us a specialized but still
exact relation, but the truncation step produces an approximate relation. This
approximation is good in a great number and variety of practical situations and problems;
we shall use (7.3) in many instances in this chapter and Chapter 8. Nonetheless, we must
always remember to ask ourselves if the approximation is warranted in the situation at
hand. To underscore the importance of this vigilance, let’s consider the process of the
supersonic jet nozzle, which consists of a convergent and divergent nozzles ganged
together to produce a discharge at supersonic velocity, as shown in Figure 7.4.

1 2 u2
Divergent ∆H = −∆
nozzle 2
Combustion Exhaust  u 22 u12 
products H H 2 − H 1 = − − 
1 uS H2
 2 2 
u1 Sonic u2
Subsonic Convergent Supersonic
nozzle
Figure 7.4:
First law applied to supersonic jet nozzle in which
the kinetic energy change is prominent

To be sure, many analyses and calculations (in other disciplines such as Fluid Mechanics)
are needed for the complete understanding, design and construction of the supersonic jet
nozzle. But to begin with, we need a first law analysis. The work-horse equation (7.3), an
approximate relation, would fail us miserably here because it leaves out the vital kinetic
term, the very term that accounts for the propulsion of the jet engine, of which the
supersonic jet nozzle is a part. We have to retreat to the step just before the truncation and
look at the complete first law expression for the steady state flow process as given by
(7.2a):

129
u2
∆H + ∆ + g∆z = Q − WS
2

The heat loss of the nozzle is insignificant ( Q → 0 ); so is the change in gravitational


potential energy ( g∆z → 0); no shaft work is transacted in the process ( WS = 0 ). So, for
the supersonic jet nozzle, the first law analysis yields:

u2
∆H = −∆
2

The propulsion comes from the change in momentum implicit in the change in velocity
from the fuel + air flow in the intake to the flow of combustion product mixture in the
exhaust. If we applied the truncated equation (7.3), we would arrive at the meaningless
result ∆H = 0. So where does the jet propulsion come from?

7.3.2 Closed Process

A closed process is one in which no transaction of material occurs between it and the
surroundings. The refrigeration process is a good example of a closed process. Therefore,
we have m 1 = m 2 = 0 . The first law expressed by equation (7.1) is reduced to:

d
(mCU C ) = mC dU C + U C dmC = q − w S
dt dt dt

Furthermore, in a closed system the material hold-up m C in it is constant, or dmC dt = 0 .


The first law expression thereby further reduces to:

dU C q w 1 dq 1 dwS
= − S = −
dt mC mC mC dt mC dt

Or:

dq dwS
dU = − = dQ − dWS
mC mC

Whence:

state 2 state 2 state 2

∫ dU = ∫ dQ − ∫ dW
state 1 state 1 state 1
S (7.4)

130
Where: Q = q mC and WS = wS mC are the heat and shaft work transactions per unit
mass of material held up in the closed system. We have also dropped the subscript C in U
without fear of ambiguity. (Note that we have not yet truncated any quantity which we
may consider insignificant; so every relation so far is exact, though greatly reduced.) We
can integrate the last expression (7.4) to obtain ∆U in its relation to the heat and work
transactions.

The internal energy U of material is known as a state function. (Other thermodynamic


quantities encountered in this chapter are also state functions: enthalpy H, Gibbs free
energy G and Helmholtz free energy A.) The differential of a state function is known as
an exact differential; therefore dU is an exact differential. For our purpose here an exact
differential such as dU is one that can be integrated between two states of the material to
yield a physically meaningful result, since the value of U depends only on the state and
not on how the state is reached. This means that we can evaluate the first integral of (7.4)
as follows:

state 2

∆U = ∫ dU = U
state 1
2 − U1

The above result means that the material has internal energy value of U 1 at state 1
and U 2 at state 2. The change in value ∆U = U 2 − U 1 is determinate, depending as it does
only on the initial state 1 and the final state 2. It is independent of how the system
evolves from the initial to the final state.

However, the same cannot be said of the integrals involving heat and work, namely the
second and third integrals of (7.4). Heat and work are not state functions. These two
integrals cannot be evaluated in the same manner as the first integral; that is to say:

state 2 state 2

∫ dQ and
state 1
∫ dW
state 1
S are indeterminate.

Why not?

Heat and work are not state functions; they are not even properties of a material
substance. Because of this nature of heat and work, dQ and dWS are not exact
differentials. Q and W S are the exchanges of heat and work between the system and the
surroundings when the material undergoes a change of state brought about by some
process, physical or chemical, or both. The values of these exchanges of heat and work
depend on the path taken by the process. All we can say for the time being is that:

∆U = U 2 − U 1 = Q ′′ − W ′′

131
Where: Q ′′ and W ′′ represent the heat and work exchanges, the values of which depend
on the particular path the process takes in bringing about the change of state in the
material. Suppose the change in material state is such that ∆U = U 2 − U 1 = 1000 kJ/kg.
This change of state can be brought about by any number of paths. Let’s consider three of
these. Path1 results in an exchanges of heat and work as (Q ′′,W ′′)Path 1 = (5000 kJ/kg, 4000
kJ/kg); similarly, (Q ′′,W ′′)Path 2 = (7500 kJ/kg, 6500 kJ/kg) and (Q ′′,W ′′)Path 3 = (1000 kJ/kg,
0 kJ/kg). Each set of (Q ′′,W ′′)Path n corresponds to the same value of ∆U = U 2 − U 1 = 1000
kJ/kg. There is any number of sets of (Q ′′,W ′′ )
Path n
just as there is any number of paths.

How do we get out of this conundrum?

Consider a closed system consisting of a material substance enclosed in a piston-cylinder


arrangement, as shown in Figure 7.5. For simplicity of presentation, the surroundings are
the ambient; the piston is frictionless; and the whole assembly is well insulated enough so
that no heat is transacted through the wall with the ambient. Initially the system is in state
1 in mechanical equilibrium with the ambient. Heat is delivered slowly into the material
via the heating coil so that the system remains at all times in mechanical equilibrium with
the ambient. The system evolves toward state 2 as heat is being absorbed by the enclosed
material and as work is being done by the rising piston against a connected load. When
state 2 is reached total heat absorbed is Q and the shaft work done is WS .

WS WS
P0 Load P0 Load

h
Heat Q (delivered by heating coil)
z2
Heat coil Heat coil
z1 Shaft work WS (done to turn a shaft
connect to a load)
State 1: (P1 ,V1 ) State 2: (P2 ,V2 )
Material properties: Material properties:
U1 , H1 U2,H2
Figure 7.5:
Evolution over a time interval of enclosed material from state 1 to state 2
entailing exchanges Q, WS

Both Q and W S are expressed on the basis of one unit mass of material held up in the
cylinder. In the material property list in Figure 7.5, we have included H . We know H is a
state function because H = U + PV by definition; here H and U are each expressed on
per mass of material held up in the closed system. U is a state function and P, V are

132
parameters defining the state. Since H is a state function, its change between two states is
independent of how this change is brought about. Because of this independence of how H
is caused to change, we apply (7.3) to the closed system as shown in Figure 7.5 as:

∆H = H 2 − H 1 = Q − WS

We are justified in applying (7.3) to the set-up in Figure 7.5 because in this set-up both
the gravitational terms (z 2 – z 1 )g and the kinetic terms (u 1 2 – u 1 2)/2g are negligible. In the
steady state flow process the change of state is due to the fact that the material leaves the
system in a state different from the one in which it enters it; the state of the material, as
well as its mass, held up inside the system remain constant with time. Here, in the closed
system, the change of state occurs in the material (constant mass) held up inside the
system as this material, in the course of system exchanging heat and work with the
surroundings, evolves over time from initial state 1 to final state 2. In the steady state
flow process, each quantity in (7.3) is expressed on the basis of unit mass of material
entering; whereas applied to a closed system, each quantity in (7.3) is expressed as unit
mass of material held up in the system. Now, since ∆H = ∆U + ∆(PV ) by definition, we
have finally our work-horse equation for the first law applied to the closed system:

∆U = ∆H − ∆PV = Q − [WS + ∆(PV )]

Or:

∆U = Q − W (7.5)

Where: W = W S + ΔPV. We shall see the distinction of W and W S below.

It should be clear that the two final expressions for the first law, ∆H = Q − WS (7.3) and
∆U = Q − W (7.5) are just alternatives. It is only as a matter of convenience that we
sometimes prefer to use H for steady state flow process and U for the closed process.
Both expressions are valid either for the open steady state flow process or for the closed
process. All we need to know is that in the steady state flow process all quantities in these
two alternative expressions are stated on the basis of unit mass of material entering;
whereas in the case of the closed process it is on the basis of unit mass of material held
up inside the system.

To appreciate the distinction between W and WS , we see that in the piston-cylinder


arrangement shown in Figure 7.5 above, we have P1 = P2 = P0 (constant ambient
pressure, since we stipulate that the system be in mechanical equilibrium with the
ambient pressure at all time). We have then:

W = WS + ∆PV = WS + P0 ∆V

133
We recognize that P0 ∆V is the work done by the piston moving against the ambient
pressure P0 ; only WS , called shaft work, is the useful part of the total work W actually
done by the system to drive a load.

To arrive at the final expression ∆U = Q − W , we have used the idealization that the
piston be frictionless and external heat be imparted to the material in an infinitesimal rate
so that the system remains at all times in equilibrium with the surroundings, i.e.
P1 = P2 = Po at all times during the evolution of the enclosed material from state 1 to state
2. This might lead some to believe that this final expression (7.5) has only a very
restricted validity. This is not true. Take note that the heat quantity Q in the final
expression is the heat absorbed by the material in the system and W the work done by
system. If the idealization conditions are removed, i.e. if we apply the analysis to a real
process, work is done against the friction and dissipated as heat QF ; and, as heat is
delivered to the system at a finite rate, the piston may advance with a jerky motion
involving acceleration and deceleration, which also results in yet more dissipated heat QM
due to the eddies in the material caused by the jerky motion. Also we note that heat
QE escapes, since no insulation is perfect. To drive this real process from the state 1 to
state 2 to deliver the same work W , a quantity of external heat QReal must be delivered by
the heating coil. When we take into account the dissipated and escaped heat, we have
then: Q = QRe al − (QF + QM − QE ) , where we have used the convention that the dissipated
heat is take as positive and the escaped heat as negative. The idealization was introduced
so that we could set Q F = Q M = Q E to zero, in order to minimize heat accounting for the
ease of presentation; it by no means restricts the validity of the final expression
∆U = Q − W to the ideal process, which states that the change in internal energy ΔU
equals to the net heat Q transferred to the system less the work W done by the system
This statement is in fact true for any process, real or ideal, as long as we are justified in
considering negligible the change in and gravitational potential energy per unit mass (z 2 –
z 1 )g and the change in kinetic energy per unit mass (u 2 2 – u 1 2)/2g of the enclosed
material.

7.4 Second Law

In the same way as we formulate the first law, we approach the second law by examining
a system (consisting of processes) in its transaction with the surroundings through the
flow of material and exchanges of heat and work, as shown in Figure 7.6. Earlier, we
showed the system as consisting of the specific process of a work-producing turbine and
its lube oil ancillary, only to take advantage of our familiarity with this particular process.
The first law and the second apply to a system consisting of any number or variety of
processes; the processes may involve physical or chemical changes or both. In Figure 7.6,
therefore, we choose not to use any specific processes in our illustration to avoid
cluttering the figure with non essential details.

134
Surroundings

System
Material m 1 1 2
Material m 2
Shaft w ork w S 1
Processes Shaft w ork w S 2
Heat q1
Heat q 2

Source Sink
at θ1 at θ2

Figure 7.6:
Second law: Entropy accounting

Referring to Figure 7.6, we list the parameters and material properties we need for the
formulation of the second law as follows:

Point 1: Point 2 Inside system


Material flow rate, kg/s: m 1 m 2
Entropy of material, J/kg K: S1 S2 SC
Velocity of material (bulk flow), m/s: u1 u2
Elevation, m: z1 z2
Material hold-up, kg: mC

Note that we indicate the absolute temperature of the heat source and sink as θ1 , θ 2 ,
respectively. We use θ instead of T to represent temperature to avoid unwieldy multiple
subscripts that would arise if we were to stick to T for temperature. We reserve T to
denote the temperature of the material substance. For the exchanges of heat, work and
material flowing we use, as before, the subscript 1 for entering and 2 for leaving the
system, and C for material held up inside the system.

7.4.1 Entropy of Heat, Work, Kinetic and Potential Energies

A quantity of heat Q available at absolute temperature θ has associated with it entropy S Q


given by:

Q
SQ =
θ

135
A quantity of work W is equivalent to the same quantity of heat available at θ → ∞ : for
example, 1000 kJ of W is equal to 1000 kW of Q available at an infinitely high
temperature. The entropy SW associated with work W is therefore zero, as:

W 
SW =   =0
 θ  θ →∞

For other forms of energy (kinetic, potential, electrical, etc), their associated entropy is
zero if they can be converted completely to work in an ideal engine. Potential and kinetic
energies meet this criterion. The entropy S P and S K associated with potential and kinetic
energies, respectively, are each equal to zero as:

 zg   u2 2 
SP =   = 0 and S K =   =0
 θ  θ →∞  θ  θ →∞

The associated entropy of the different forms of energy (heat, work, potential, kinetic,
electrical, chemical potential, etc) lies in the very core of the nature of entropy. To
examine it any further in the confine of this short overview would take us too far afield.

7.4.2 Formulation of the Second Law of Thermodynamics

The second law says that:

A system (consisting of processes) interacts with the surroundings through the exchanges
of material substance, heat and work (work including kinetic and potential energies
transacted by the material flow). This interaction, for real processes, always results in a
net increase in entropy, with the system and surroundings taken as a whole. For an
ideal (or reversible) process this net increase in entropy is zero.

So, let’s do the entropy accounting:


Entropy Entropy
increase decrease
in system in surroundings
Due to:
Material flow, m 1 , m 2 m 2 S 2 − m 1 S1 ------
d
Material accumulation in system, mC (mC S C ) ------
dt
q1 q 2
Heat exchanges, q1 , q 2 ------ −
θ1 θ2
Work exchanges, w S 1 , w S 2 ------ 0

136
u12 u 22
Material kinetic energy, m 1 , m 2 0 ------
2 2
(bulk flow)

Material potential energy, m 1 z1 g , m 2 z 2 g 0 ------

We then have the expression of the second law as:


(m 2 S 2 − m 1 S1 ) + d (mC S C ) ≥ q1 − q 2
 
(7.6)
dt θ1 θ 2

This complete expression can be specialized to the steady state flow process and the
closed process.

7.4.3 Steady state flow process

At steady state:

m 1 = m 2 = m (Material balance)

d
(mC S C ) = S C dmC + mC dS C = 0
dt dt dt

Since,
dmC dS
= 0 and C = 0 (No material accumulation and no change in state with time).
dt dt

The expression (7.5) reduces to our work-horse equation:

1  q1 q 2  Q1 Q2
S 2 − S1 = ∆S ≥  − = −
m  θ1 θ 2  θ1 θ 2
Or:
Q1 Q2
∆S ≥ − (7.7)
θ1 θ2

Where: Q1 = q1 m 1 and Q2 = q 2 m are each the heat transacted per unit mass of material
entering the system.

7.4.4 Closed Process

For closed process, we have from (7.6) with m 1 = m 2 = 0 and mC = constant:

137
d
(mC S C ) ≥ q1 − q 2 ; or:
 
dt θ1 θ 2

SC 2
1 
state 2
dq1
state 2
dq 2  1  1
state 2
1
state 2


∫ dS C = S C 2 − S C1 = ∆S C ≥ ∫
mC  state 1 θ1 dt
dt − ∫ dt = ∫ dq1 −
θ dt  mC  θ1 state 1 θ2 ∫ dq 2

S C 1! state 1 2 state 1 

The first integral involves entropy, which is a state function and dS C is therefore an exact
differential; so it can be evaluated, independent of the path the process may take to
evolve from state 1 to state 2. As encountered in the formulation of the first law, we are
faced again with the same conundrum of dq1 and dq 2 being not exact differentials. They
cannot be integrated to give physically meaningful results in the same manner as dS C is
integrated to obtain the meaningful result ΔS C = S 2 – S 1 . All we can say is that the
change in material entropy ∆S C arising from the change of state of the material held inside
the system is tentatively related by:

Q1′′ Q2′′
∆SC ≥ −
θ1 θ2

state 2 state 2
1 1
Where: Q1′′ ≠
mC ∫ dq1 and Q2′′ ≠
state 1
mC ∫ dq
state 1
2 , but are yet to be identified.

Here, however, we have recourse to the final work-horse expression (7.7) for the open
steady state flow process. There, the change in entropy of the material ∆S is brought
about by the fact that the material enters and leaves the system in different states, hence
different entropy; while the state of the material and its mass held up inside the system
are constant in time, i.e. no accumulation. This ∆S is related to the heat exchanges as
shown in (7.7). Here, in the closed process, the change in material entropy ∆S C is brought
about by the change of state over time of the material held up inside the system, while
there is no material entering and leaving. We can see this by examining the unreduced
Second Law (7.5), as:
d
(mC S C ) ≥ q1 − q 2
 
For steady state flow process: (m 2 S 2 − m 1 S1 ) +
dt θ1 θ 2

d
(mC S C ) ≥ q1 − q 2
 
For closed process: (m 2 S 2 − m 1 S1 ) +
dt θ1 θ 2

Therefore we can say for the closed process, dropping the subscript C on S without
incurring ambiguity:

138
Q1 Q2
∆S ≥ − (7.7a)
θ1 θ 2
Where: Q1 ,Q2 (= q1 mC , q 2 mC , respectively) are each the heat transacted (over the time
interval of the evolution of state) per unit mass of material hold-up in the system; they
identify with the erstwhile undetermined Q1′′ and Q2′′ used in our tentative expression
above.

We see here the final expression of the second law is the same for the closed process as it
is for the open steady state flow process. It should be noted that, in either case, we say the
increase in entropy ∆S of the material (due to the change of state of the material) is
greater than the net decrease in entropy of the surroundings (due to heat transaction
between the system and the surroundings). The only difference between (7.7) and (7.7a)
is that, in the former (open steady state flow process) all quantities are stated on the basis
of unit mass of material entering the system; whereas in the latter (closed system) all
quantities are on the basis of unit mass of material held-up in the system.

7.5 Differential Forms of the First Law

We have two expressions for the first law: ∆H = Q − WS (7.3) and ∆U = Q − W (7.5).
These expressions are said to be in the finite forms because we deal with the finite
difference ∆ of the quantities U and H as ∆U and ∆H . However, in many analyses we
also need the differential forms of the first law; in other words we need expressions
for dU and dH instead of ∆U and ∆H . Let’s first state these differential forms and
subsequently show how they come about:

dU = TdS − PdV Or, alternatively: (7.8)

dH = TdS + VdP (7.8a)

These two expressions, (7.8) and (7.8a) are customarily called the differential forms of
the first law; although, strictly speaking, they are the results of combining the first and
the second law, as we will see in a moment.

To see how these differential forms are obtained, let’s consider the idealized closed
system shown in Figure 7.7. The piston is frictionless. Since the system is closed it has no
material transaction with the surroundings, but only transaction in heat and work.
Initially, the system at pressure P, with the total weight of the piston assembly Mg, is in
equilibrium with the surrounding pressure P o . Suppose infinitesimal amounts of heat are
transacted so that the system gains an amount of heat δQ = δQ1 − δQ2 . This gain in heat
brings about a change in state of the enclosed material measured by dU and dS ; it also
causes the enclosed material to expand by an infinitesimal amount dx. . We can apply the
first and the second laws, equations (7.5) and (7.7a), to this situation as:

139
dU = δQ − δW (7.9)
δQ δQ
dS ≥ 1 − 2 (7.10)
θ1 θ2

Surroundings
System δQ = δQ1 − δQ2
M
δW dU = δQ − δW
P0
As θ1 → T → θ 2 :
dx δQ2
TdS = δQ ;
Piston area = A P, V , T
PdV = δW and
δQ1 U , S, H Sink
Source
at θ 2
dU = TdS − PdV
at θ 1

Figure 7.7:
Differential form of the first law

As the enclosed material expands by dx, it does work in the amount PAdx = PdV, where
A is the area of the piston. However, the manifested work done is δW , the work done in
lifting the piston assembly of total mass M and in pushing the piston against the
surrounding pressure P o . Even though we work with the premise of a frictionless piston,
the expansion may still be “jerky”; this is so because heat transfer must occur at a finite
rate, which in turn causes instantaneous thermal disequilibrium between the system and
the surroundings. The piston may advance in a series of accelerations and decelerations
before it comes to rest in equilibrium with the surrounding, because, in the course of its
expansion, the system pressure may momentarily exceed its equilibrium value. This jerky
motion is dissipative, hence irreversible, and does not contribute to the manifested work
done δW . This said, we can state that:

δW ≤ PdV (7.11)

The equality holds if the expansion is reversible (i.e. without jerky dissipative motion).
We can idealize a reversible process for the transactions of heat and work by assuming no
friction and letting the heat source temperature θ1 approach the system temperature T, and
T in turn approach the heat sink temperature θ 2 , i.e. θ1 → T → θ 2 . With this limiting
process the system is then at any instant in thermal and mechanical equilibrium with the
surroundings. That is to say: the system temperature T is in equilibrium with the
temperatures of the heat source and heat sink; and the system pressure P is in equilibrium
with the surrounding pressure P o . The process is reversible and we can now state,
invoking the equality in the expressions of (7.10) and (7.11):

140
δQ1 δQ2 δQ − δQ2 δQ
As θ1 → T → θ 2 dS ≥ − → dS = 1 = ; and
θ1 θ2 T T
δW ≤ PdV → δW = PdV

Substituting these reversible expressions of δQ and δW into the first law (7.9), we have
the differential form of the first law as: dU = TdS − PdV . Using the definition
H = U + PV (i.e. dH = dU + PdV + VdP ), we obtain the alternative form:
dH = TdS + VdP . These two differential forms are shown as (7.8) and (7.8a) above.

It is commonly believed that these differential forms of the first law are valid only for
reversible processes, because, manifestly, we have invoked the idealized condition of
reversibility in arriving at them. This common belief is mistaken. We see that in the
relations dU = TdS − PdV and dH = TdS + VdP , the differentials are all those of state
functions and state parameters; they are therefore exact differentials. To describe a
change from state 1 to state 2, we only have to integrate these expressions between state 1
and state 2:

State 2 State 2

∆U = U 2 − U 1 = ∫ TdS −
State 1
∫ PdV ; and
State 1

State 2 State 2

∆H = H 2 − H 1 = ∫ TdS +
State 1
∫ VdP
State 1

These integrations are all independent of any path by which the evolution from state 1 to
state 2 may occur, since the quantities S appearing in the integrals is a state function and
P, V, T are parameters defining a state. Therefore these differential forms of the first law
are valid for all processes, reversible or not.

As another way to look at it, we can argue that both dU = δQ − δW and dU = TdS − PdV
are valid for all processes, reversible or not. It just so happens that for a reversible
process we have δQ = TdS and δW = pdV , which is demonstrated by invoking an
idealized reversible process as we did above; otherwise δQ < TdS and δW < pdV . Or, in
their integrated forms, we can say:

State 2 State 2

Q≤ ∫ TdS and W ≤
State 1
∫ pdV , where the equality holds for a reversible process; so that:
State 1
State 2 State 2

∆U = U 2 − U 1 = Q − W = ∫ TdS − ∫ PdV is valid for all processes, reversible or not.


State 1 State 1

We have discussed this at some length in the hope of dispelling any misconception about
the validity of the differential forms of the first law. We find that misconceptions of this

141
nature could be a serious impediment to the understanding and application of
thermodynamics; they can in fact lead a person astray for a long time in his attempt to
understand the basic principles of thermodynamics and their applications to engineering.

7.6 Entropy and Internal Energy: the Fundamental Duo (U , S )

Internal energy U

Internal energy U of a material is a fundamental property. For a pure substance, it is


uniquely determined by the specification of any two of the parameters P , T and V; i.e.
U = U (P, T ) , or U = U (T ,V ) , since V = V (T , P ) . For a mixture of substances, we
have U = U (P, T , m j , x j ) or U = U (T , V j , m j , x j ); where m1 , m2 ...m j .... are the pure
components making up the mixture and x1 , x 2 ...x j .... are the corresponding composition
and V1 , V2 ...V j .... are the specific volumes of the components as they exist in the mixture.

Internal energy is the energy contained in the material by virtue of its molecular and
nuclear structures and the interactions among the molecules (intermolecular forces) in the
material at a given pressure and temperature. U is truly the energy content of the
material; and this content can be determined. Conceptually, we can evaluate at a
fundamental level the internal energy of a material U by considering:

• Kinetic energy of the molecules by virtue of their translational and rotational


motions, as well as the energy of their vibrational motions;
• Energy levels of the electrons orbiting the atoms of the molecules;
• Intermolecular forces; and
• Constitution of the nuclei.

The kinetic energy referred to in the first consideration is distinct from the kinetic energy
term u 2 2 g due to the bulk flow at macroscopic scale of the material considered in the
formulation of the first and the second laws. The kinetic energy u 2 2 g due to the bulk
motion of the material is not part of its internal energy U. The last three considerations
result in what is generally called binding energies. The fourth consideration is included
for the sake of completeness. It comes into play only if the process entails nuclear
transformation. In most applications in process engineering, however, the nuclei remain
intact throughout and only the re-arrangement of the valence electrons (chemical
transformation) and or intermolecular forces (physical transformation) need be
considered.

Entropy S

Entropy S of a material substance, like internal energy U , is a fundamental property.


Entropy is statistical in nature. Amazingly, at the fundamental level entropy is expressed
in a supremely simple relation (due to Ludwig Boltzmann):

142
S = k ln W ; or (7.12)

S = R ln W (7.12a)

Where: k is the Boltzmann constant, R the universal gas constant and W the number of
distinguishable ways the molecules can be arranged for a given total internal energy U .
Note: S in (7.12) is expressed on the basis of per molecule (or per particle) and S in
(7.12a) on the basis of per unit mole, with R = N A k , N A being the Avogadro’s number.
W does not stand for work (as in U = Q − W ). The value of W can be obtained from
classical statistics, also known as the Maxwell-Boltzmann statistics; or, usually in cases
of extremely low temperatures, from one of the two quantum statistics: the Fermi-Dirac
statistics for molecules (or particles) behaving like fermions, or the Bose-Einstein
statistics for molecules (or particles) behaving like bosons. The two quantum statistics
approach the classical as temperature rises.

However, in process engineering and in most other applications, seldom do we have to


resort to Boltzmann’s fundamental relation (7.12) to find the entropy of a material
substance; instead, we calculate it, as well as other useful thermodynamic quantities,
from the macroscopically measurable properties of the material, namely its PVT data and
its specific heats C P and C V . This will be treated in the subsequent sections.

7.7 The Power of Statistics

Any macroscopic sample of material contains an enormous number of individual


molecules, in the order of 1023. Thanks to the large number, the full power of statistics is
brought to bear. This means we don’t really have to measure the binding energies held in
individual molecules or the intermolecular forces among them; nor do we need to
calculate entropy, although we can, from the number of distinguishable ways the
molecules can distribute themselves to yield a given total energy of a macroscopic
sample of the material. Phew! More importantly, this means that we can evaluate the
internal energy and entropy, U and S , from measurable and familiar macroscopic
properties of the material. Once U and S are known, all other thermodynamic quantities
of like H and others can be calculated, because they are all defined on U and S.

Take a macroscopic sample of material captured in a container or in a “snapshot” of a


body of flowing material. If the material is a gas, the molecules move about randomly
with a wide range of velocity, colliding at random with each other and with the container
wall; they also rotate and vibrate. Molecules in a liquid move less extensively; they rotate
and vibrate. Molecules in a solid vibrate. These various random motions and collisions
give rise to a range of kinetic energy, momentum, intermolecular distance, and
intermolecular force. These dynamical quantities of the ensemble of molecules each have
their macroscopic manifestation as follows:

143
Dynamics at microscopic level Macroscopic manifestation

Average kinetic energy (translation and rotation)


and energy of vibration Temperature, T

Average magnitude of linear momentum Pressure, P

Average intermolecular distance (mean free path) Specific Volume, V

Average kinetic energy (translation and rotation),


energy of molecular vibration and intermolecular force Specific heat, CV , C P

The macroscopic parameters and properties: P,V , T , CV and C P are all measurable and are
in essence statistical averages of the underlying microscopic dynamics. Consequently, the
values of internal energy and entropy, calculated as they are from P,V , T , CV and, C P , are
themselves statistical averages. Given such an enormous number of molecules involved,
any significant or palpable deviation from the average is so highly improbable that we
can, to all intents and purposes, take following expressions for U (T , V ) , U (T , P ) ,
S (T , V ) and S (T , P ) , which are in essence statistical averages, to be deterministic. Such is
the power of statistics when applied to an ensemble of a large number of particles or
molecules!

Specifically, we can calculate U and S from the macroscopic measurable quantities as:

T V
  ∂P  
U (T , V ) − U (T0 , V0 ) = ∫ CV dT + ∫ T   − P  dV ; or, alternatively, (7.13)
To V0  
∂T V 
T P
  ∂V     ∂V   ∂V  
U (T , P ) − U (T0 , P0 ) = ∫ C P − P   dT − ∫ T   + P  dP (7.14)
T0   ∂T  P  P0  
∂T  P  ∂P  T 
T V
CV  ∂P 
S (T , V ) − S (T0 , V0 ) = ∫ dT + ∫   dV ; or, alternatively, (7.15)
T0
T V0 
∂T V
T P
CP  ∂V 
S (T , P ) − S (T0 , P0 ) = ∫ dT − ∫   dP (7.16)
T0
T P0 
∂T  P

Where: CV = (∂U ∂T )V and C P = (∂H ∂T )P are the specific heat of the material at
constant volume and constant pressure, respectively; P0 , T0 are reference pressure and
temperature, V0 being the specific volume corresponding to P0 , T0 . At this reference
condition (P0 , T0 , V0 ) of our choice, the internal energy and entropy, namely
U 0 = U (T0 ,V0 ) = U (T0 , P0 ) and S 0 = S (T0 ,V0 ) = S (T0 , P0 ) are assigned an arbitrary
constant value of usually but not necessarily zero. The above expressions for U and S,

144
equations (7.13) to (7.16), are derived from the differential forms of the first law, (7.8)
and (7.8a), and a series of relations known as the Maxwell relations. The derivations of
these expressions are shown in Appendix VIII.

It should be noted, since the fundamental duo (U , S ) can be calculated from the measured
quantities pertaining to the material, so can the derived state functions such as H and
others can be calculated in a similar way, derived as they are from the fundamental duo.

The two expressions U (T , V ) and U (T , P ) , in (7.13) and (7.14), are alternative to each
other. Suppose we want to find the internal energy given T , P we can use the expression
for U (T , P ) and perform the required integration. But, given T , P , we can find V from the
equation of state V = V (T , P ) ; so that with the duo (T , V ) we can use the
expression U (T , V ) . In essence, only one of these expressions suffices. However, one of
the expressions may be less laborious to use than the other, depending on how the
measured PVT data are organized or presented, or in other words, on the forms the
equations of state, V = V (T, P) or P = P (T, V), constructed from the measured PVT data.
So, the use of one alternative over the other in a particular situation is just a matter of
convenience. The same can be said of the pairs S (T, V) and S (T, P)] in (7.15) and (7.16).

It is seen that the functions required for the evaluation of the integrals in equations (7.13)
to (7.16) are:

CV = CV (T , V )
C P = C P (T , P )
 ∂P 
  = f1 (T , V ) , or a function of T and V
 ∂T V
 ∂V 
  = f 2 (T , P ) , or a function of T and P
 ∂T  P
 ∂V 
  = f 3 (T , P ) , or a function of T and P
 ∂P  T

The first two functions, CV (T , V ) and C P (T , P ) , are constructed using data from
calorimetric measurements. The last three functions f1 (T , V ) , f 2 (T , P ) and f 3 (T , P ) can be
obtained from equations of state constructed from measured PVT data organized in the
forms of P = P(T ,V ) and V = V (T , P ) . Note that CV , C P are related to each other through
the equations of state as:

2
 ∂V   ∂P   ∂V   ∂P 
C P − CV = T     = −T    
 ∂T  P  ∂T V  ∂T  P  ∂V  T
This relation makes it possible to use the equations of state to check the consistency of
CV , C P obtained from calorimetric measurements. It is easily seen from this relation that

145
for solids and liquids C P ≅ CV , since (∂V ∂T )P is small; therefore for solids and liquids,
we only talk about one specific heat.

Note that the above expressions for U (T , V ) , U ( P, T ) , S (T , V ) and S (T , P ) do not include


changes in U and S brought about by material transformation, chemical or nuclear. For a
process involving chemical transformation (or chemical reactions) different molecules
are produced by the re-arrangement of the valence electrons. This re-arrangement is
accompanied by changes in internal energy and entropy, ∆U R and ∆S R , which are given
by:

   
∆U R =  ∑ xi U i ( products) + I P  −  ∑ y jU j ( reactants) + I R 
 i   j 
   
∆S R =  ∑ xi S i ( products) + J P  −  ∑ y j S j ( reactants) + J R 
 i   j 

Where: U i , U j , S i , S j are each given by the expressions for pure components U = U (T , V ) ,


U = U (T , P ) , S = S (T , V ) or S = S (T , P ) in equations (7.13) to (7.16); x i are mole
fractions of the products and y i are mole fractions of the reactants. Implicit in the
summations ∑ xiU i , etc, is that the mixtures of products and reactants are treated as
i
ideal mixtures. In recognition that these mixtures are real, I P , I R , J P and J R are
corrections added to the respective sums. The corrections for non-ideality, in our context
here and in general, form a vast and evolving area of study and research in applied
thermodynamics. The expressions of these corrections are mostly empirical; they form
the cornerstones of a thermodynamic data base. It is beyond the scope of this short
overview to treat them any further than just introducing them here as concepts.

7.8 Derived Quantities

For all material substances, there are two and only two fundamental quantities in
thermodynamics. They are internal energy and entropy: U, S. They are fundamental
because they have direct physical correspondence with the microscopic dynamics of the
material, as we discussed above, and they can be calculated at a fundamental level from
these dynamics. They are also state functions, so that, given its ( U , S ), the state of a
material substance is completely defined. Nothing more is needed. However, in
thermodynamics, we also work with a few derived quantities. These derived quantities
come about solely as a measure of convenience. They are non-fundamental and, though
convenient and useful, they are non-essential. For example, if the steam table, or any
other data base for that matter, shows only U and S, then we have the complete data set
that would enable us to do any required thermodynamic calculations.

146
We have already met the derived quantity H , named enthalpy, in the formulation of the
first law; it is convenient to use in the analysis of an open steady state process. There are
two more named derived quantities used often in process engineering, G and A . We list
them all here:

Enthalpy: H = U + PV
Gibbs free energy: G = (U + PV ) − TS = H − TS
Helmholtz free energy: A = U − TS

Consequently, we have G = A + PV as we have H = U + PV . So you see, these derived


quantities just move in a circle centered on the fundamental state functions U and S and
state parameters P, V , T . As a matter of fact, more derived quantities, besides H, G and A,
can be generated by the use of a mathematical process called the Legendre transform.
(Some details on the use of the Legendre transform can be found in some texts: Callen(8);
Boas(9)). We would then produce more derived quantities going in a circle, telling us
nothing more about the material state that is already completely determined by the
fundamental duo: (U , S ) .

Gibbs free energy (or Gibbs function) G and Helmholtz free energy (or Helmholtz
function) A arise from the analyses of chemical and phase equilibria. In these analyses,
the clusters (U + PV ) − TS (for a flow process) and U − TS (for a closed process) pop up,
in the manner that the cluster U + PV pops up in the formulation of the first law. We gave
each of these clusters a symbol and a name. That is all there is to it for H , G, A . This
measure of convenience turns out to be so useful that the values of H , G are frequently
given directly in many data bases, just to spare us the labour of calculating them from the
fundamental duo (U , S ) . Apart from the obvious advantage of using a single symbol for a
cluster of them, these derived quantities are also convenient for the narrative. For
examples, when a change of state is accompanied by no exchanges of heat and work, we
call it isenthalpic or ∆H = 0 ; phase or chemical equilibrium is stated by saying ∆G = 0
or ∆A = 0 . Conversely, for example, if we want to know how things would shake out if a
system were to achieve equilibrium, we can impose the requirement ∆G = 0 and look at
the results that must follow as a consequence of this imposition. To underscore the non-
essential nature of these derived quantities, the afore-mentioned situations can be stated,
rather clumsily, ∆(U + PV ) = 0 , ∆(U + PV − TS ) = 0 , ∆(U − TS ) = 0 instead of ΔH =0 ,
ΔG =0 , ΔA=0

All the derived quantities are also state functions, since U and S, from which they are
derived, are state functions. The derived quantities, like the fundamental duo (U , S ) , can
therefore be calculated from the physically measurable quantities and parameters of the
material substance, namely: P, V , T , CV , C P ,

If the derived quantities like G, A had not made their appearance through the study of
equilibria, the Legendre transform would have inevitably generated them, (plus other not

147
so useful ones). But of course it was their appearance in equilibrium study that inspired
the use of the Legendre transform in retrospect.

7.9 Third Law and Absolute Zero Entropy

There is a “curious” little known law called the third law of thermodynamics. It arises
from the application of quantum statistics at low temperatures. The third law states: At
absolute temperature T = 0 , S = 0 and U > 0 . It can be seen from S = k ln W (7.12) that,
at absolute zero, the molecules can only arrange themselves in only one way
( W = 1 ; ln 1 = 0 ). So entropy does have an absolute zero, by its very statistical nature!
The fact that energy retains a positive absolute minimum value at T = 0 is due to two
quantum phenomena: the Pauli Exclusion Principle and the Heisenberg Uncertainty
Principle. The energy of a material cannot be zero, in the absolute sense.

In applications involving energy and entropy (U, S) and derived quantities such as H, G
and A. we are invariably interested only in their changes and not their absolute values.
That’s why the choice of reference zero for energy and entropy is arbitrary. We have
similar experience with temperature, where we have two common scales: Celsius and
Fahrenheit, each with its own arbitrary reference zero. The fact that there is a true
absolute zero for temperature is of no consequence to us when we only deal with
temperature difference.

7.10 Reference State

Having introduced the derived quantities, we should elaborate the choice of the reference
in our thermodynamic data bases. In the expressions for the fundamental duo (U, S),
equations (7.13) to (7.16), we have U 0 = U (T0 , V0 ) = U (T0 , P0 ) and
S 0 = S (T0 , V0 ) = S (T0 , P0 ) , which we can set to any arbitrary constant value. The
arbitrariness is two-fold. The first lies obviously in our arbitrary choice of (P 0 , V0 , T0 ) .
Secondly, note that U 0 , S 0 are arbitrary constants of integration; any constants will do, but
we customarily set them to zero. Note that we have only two degrees of freedom to set
our arbitrary constant, due to the fact that we have only two fundamental state functions,
namely the duo (U, S). For example, if we choose to set U 0 , S 0 each to some arbitrary
constant, then H 0 , G0 , A0 cannot be set arbitrarily; they have to be H 0 = U 0 + P0V0 ,
G0 = U 0 + P0V0 − T0 S 0 and A0 = U 0 − T0 S 0 . Or, if we choose to set the pair H 0 , S 0 to
some arbitrary constants, then U 0 , G0 , A0 cannot be arbitrary but are bound by their
relations with H 0 , S 0 . Note also that we cannot simultaneously set the pair H 0 ,U 0 to
arbitrary constants, nor can we do so with the pair G0 , A0 .

This elaboration on the reference state is made in recognition of the fact the compilers of
the thermodynamic data, in commercial software or in data books, truly take to heart the

148
arbitrariness of arbitrary constants. We must remain vigilant with the sources of
thermodynamic data we use. The same entropy, for example, will have different values in
different sources, because the different data sources use different reference zeros. It is
important to remember that when we need to use data from different sources in a given
problem, we must adjust them to a common (again, arbitrary) reference zero before we
can use them.

7.11 Difficulty in Working with Thermodynamics

Thermodynamics is a cornerstone of process engineering. But it is a difficult subject.


The difficulty arises from:

1. Not understanding that there are only two fundamental state functions of a
material substance, namely U and S. In fact they are the only ones that need to be
understood. The other state functions, like H , G and A , are derived from them.
They have, therefore, no independent physical meanings. Much failure to
understand and to use thermodynamics correctly arises from erroneous and
misleading meanings ascribed to these derived quantities, meanings that they do
not have;

2. Not recognizing the limited applicability of the reduced and truncated (i.e.
specialized) expressions used in analyses and calculations; and

3. Not recognizing that heat and work are not intrinsic properties of a material
substance, but are exchanges of energy.

By and large the difficulty is dealt with by skirting around or by entertaining some
comforting notions of the thermodynamic quantities – notions that are either incomplete,
or mistaken.

The process design of fractionation, heat exchangers and fired heaters, to cite just a few,
is based on the laws of thermodynamics. Yet we can spend a lifetime doing this process
design work without the need to know or understand the thermodynamic principles on
which the work is based. We do the work by following calculation procedures already
established by others. By and large we get it right if the procedures we follow are right.
Getting it right is of course important; but even more important is knowing why it is
right. Otherwise we find ourselves in a treacherous terrain when we have to do work for
which there are no prescribed procedures to follow.

In heat exchanger design we know that ∆H = Q , because the procedure says so and
because that’s the first law; or rather, a very specialized, therefore restricted, expression
of the first law. Suppose we need to heat a material stream from state 1 (P1 ,T1 ) to state
2 (P2 ,T2 ) , using a heat exchanger. We find the enthalpy of the material H 1 at state 1
and H 2 at state 2. There we have it: the heat required Q is given by Q = ∆H = H 2 − H 1 .

149
We get it right because the procedure we follow is right; and that’s good. The bad part is
that we also imbue enthalpy with a meaning it does not have: we say enthalpy is the “heat
content” of a material. We entertain the notion that at a given state the material
“contains” an amount of heat; to change its state we either add or remove heat from it;
and we hang our hat on this notion. We are pleased with this because we have the
comfortable feeling that we have finally captured the essence of this otherwise
confounded thing called enthalpy. But in another familiar situation, as exemplified by the
steam turbine, our comfortable notion of enthalpy appears to be shaken. In the case of the
turbine (see Figure 7.8), we have ∆H = WS . So, how does the notion of heat content fit
here? We modify our notion by saying enthalpy is still heat content; but, in the case of a
work-producing turbine, the change in heat content of the material manifests itself as
shaft work W S . If this modified notion still doesn’t sit well, we extend it by saying that
the enthalpy of a material is its energy content; and a change in this energy content, i.e. a
change in enthalpy according to this notion, can manifest itself either as work or as heat,
or partly as work and partly as heat at the same time, as established by the first
law ∆H = Q − WS . In an adiabatic process, i.e. Q = 0, this energy content is converted
entirely to work. By following this line of argument, apparently very logical, we are
digging ourselves into one deep hole of misconception, as we shall see in the next
section.

Surroundings

System 2
Steam m
Turbine Shaft w ork w S
Steam m 1
Adiabatic q = 0

m (H 2 − H 1 ) = q + w S

Figure 7.8:
∆H = H 2 − H 1 = WS
First law: work-producing turbine Where: WS = w S m

7.12 Entropy, Exergy and the Degradation of Energy

To see how such mistaken notions can be treacherous, let’s examine another familiar
process of a steam letdown station, where HP (high pressure) steam is throttled down to
LP (low pressure) steam. See Figure 7.9. We see here that the notion of enthalpy as the
“energy content” of a material becomes untenable. We note that enthalpy remains
unchanged through the throttling process: H 2 = H 1 , true to the first law, because no shaft
work or heat is transacted in the throttling process. According to this notion, then, steam
has the same energy content (i.e. enthalpy) before and after the throttling process, and
none has been lost. This is counter-intuitive, we protest. At point 1 we have HP steam
and at point 2 LP steam. We know that HP steam (4420 kPa) “packs more punch” than
LP steam (446 kPa). We know instinctively and intuitively that for a given mass we can

150
generate more shaft work with HP steam than with LP steam, using, for example, a
turbine exhausting to the ambient, even though HP and LP steams have exactly the same
enthalpy of 3148.9 kJ/kg, or exactly the same “energy content” according to this
mistaken notion of energy. Our instinct is indeed correct, as shown in the simple
illustration in Figure 7.10, where the HP and LP steams in question are each made to
produce shaft work by means of a turbine operating and exhausting to same pressure sink
of 101.3 kPa (a). True to our intuition, the HP steam does indeed produce more shaft
work, quite a bit more, than the LP steam: 206.6 kW versus 97.7 kW.

HP steam header

Steam m q = 0 (adiabatic)
1
w S = 0 (no shaft work)
q w S
2 m (H 2 − H 1 ) = q + w S
BFW
∆H = H 2 − H 1 = 0
LP steam header

Figure 7.9:
First law for isenthalpic steam letdown (a.k.a. isenthalpic throttling)

We list in Table 7.1 the thermodynamic states and parameters for the steam letdown
process:

Table 7.1: States of HP and LP Steam

Pressure Specific Tempera- Enthalpy Internal Entropy Exergy


volume ture energy
P, kPa V, m3/kg T, K H, kJ/kg U, kJ/kg S, kJ/kg K E H , kJ/kg
Point 1 4420 0.0628 648 3148.9 2871.1 6.6294 1241.1
Point 2 446 0.6290 613 3148.9 2868.3 7.6561 945.5
Δ 0.0 -2.8 1.0267 -295.6

− T0 ∆S = -295.6
Reference state: S 0 = 0.2243 kJ/kg K (saturated water at T0 )
H 0 = 63.0 kJ/kg (saturated water at T0 )
P0 ,T0 = 101.3 kPa, 15 °C

If we turn to the internal energy U , which is indeed the true energy content of steam, we
see only a slight decrease (by 2.8 kJ/kg), as shown in Table 7.1. This slight decrease
in U is due to a lower temperature brought about by the Joule-Thompson effect of the
throttling process, which effect reflects the change in intermolecular potential energy due

151
to the larger intermolecular distance at the lower pressure. This decrease is far from
enough to account for the far greater work-producing capacity of the HP steam, as we
have seen. As a matter of fact, it does not account for the work-producing capacity of
steam at all, as we shall see below. So, the high school definition of energy as the
capacity to do work does not sit quite well, either.

Steam exhaust Steam exhaust


H 2 = 2405.1 kJ/kg H2 = 2797.3 kJ/kg
P 2 = 101.3 kPa (a) P2 = 101.3 kPa (a)

HP steam inlet T 2 = 373 K T2 = 433 K

H 1 = 3148.9 kJ/kg
m = 1000 kg/h LP steam inlet m = 1000 kg/h
H 1 = 3148.9 kJ/kg
P 1 = 4420 kPa (a) P 1 = 446 kPa (a)
T 1 = 648 K w S T 1 = 613 K w S
m = 1000 kg/h m = 1000 kg/h
Adiabatic efficiency = 100 % Adiabatic efficiency = 100 %
WS = ∆H = H 2 − H 1 WS = ∆H = H 2 − H 1
w S = m∆
 H = 206.6 kW w S = m∆
 H = 97.7 kW
(a) HP Steam turbine exhauting to atmosphere (b) LP Steam turbine exhauting to atmosphere

Figure 7.10:
Comparison of work producing ability of HP and LP Steam of equal enthalhpy

There is something about the quality of energy that the notion of H (or U) as “energy
content” does not capture. This is where the second law comes in. We see that the
throttling process (steam letdown) results in an increase in entropy. (See Table 7.1.) The
HP steam has the state (H , S ) = (3148.9, 6.6294); LP steam (H , S ) = (3148.9, 7.6561).
This tells us that for a given H (or U ), the state with a lower S (i.e. HP steam) has a
greater work-producing capacity. Therefore, entropy is a measure of the quality of
energy. To quantify this measure, we have introduced another derived quantity E H called
exergy. (Incidentally, E H cannot be generated by the use of the Legendre transform.) It is
defined as:

E H = (H − H 0 ) − T0 (S − S 0 ) (7.17)

Where: H 0 , S 0 are the values of H , S at the reference condition (P 0 , T 0 ), the reference


condition for the purpose of this definition being the surroundings. Given a material
substance at state (H , S ) , or equivalently (U , S ) , its exergy E H is the true measure of its
work-producing capacity. Exergy is in fact the maximum work that can be produced by
bringing the material from a given state into a state in which the material is in complete
mechanical and chemical equilibrium with the surroundings at (P 0 , T 0, S 0 ). (Exergy is
called availability in some texts.)

152
For the throttling process in Figure 7.9, we have HP steam at (H , E H ) = (3148.9, 1241.1)
and LP steam at (H , E H ) = (3148.9, 945.5), an exergy change ∆E H of -295.6 kJ/kg
(reduction). The throttling process causes a loss of work-producing capacity of steam by
295.6 kJ/kg; or, 295.6 kJ/kg of the work-producing capacity of the original HP steam is
lost, irreversibly, through the throttling process. In the throttling process, as in any real
process, energy is always conserved but also always degraded. The degradation is given,
quantitatively, by the exergy loss.

It is easy to see how entropy is a measure of energy degradation. We have from (7.17):

∆E H = E H 2 − E H 1 = (H 2 − H 1 ) − T0 (S 2 − S1 ) = ∆H − T0 ∆S

With ∆H = 0 for this particular process (isenthalpic throttling), we have ∆E H = −T0 ∆S .


This is borne out by the numerical example of the throttling process given above (Figure
7.9; Table 7.1), where ∆E H = −T0 ∆S = −295.6 kJ/kg.

This again illustrates that enthalpy is neither heat nor the potential to do work: it is
enthalpy H = U + PV. Depending on the process, a change in enthalpy can lead to a
transaction of heat (e.g. in an exchanger) or work (e.g. in a turbine), or to a combination
of heat and work transaction (e.g. in a compressor with interstage cooling); or we can
have a process entailing no change in enthalpy (e.g. in a throttling valve). In all these
processes it is exergy, not enthalpy, that tells us how much work can be realized.

7.13 Cookbook Recipe

Pressure P = 475.8 kPa (a)


Surroundings m 2 , H 2 Temperature T = 150 o C
2 w S =0
Vapour Liquid Point 2
System
q 2 = 0 Enthalpy, kJ/kg: HV HL H 2 (= H V )
P, T
Sat'd steam
Internal energy, kJ/kg: U
V UL U 2 (= U V )
q1 m 1 = 0
V2 (= VV )
3
Sat'd w ater
Specific volume, m / kg: VV VL
1
H1 Hold-up, kg: mV mL
Steam production, kg/s : m 2 = 20
Figure 7.11:
Total hold-up in system, kg: mC = mV + m L
Making steam in batch boiler

153
When confronted with problems in thermodynamics that lie outside the purview of
prescribed calculation procedures, we tend to look into the texts, in a cookbook fashion,
for equations or formulas that might meet our need. More often than not, the equations or
formulas we find are reduced, truncated and so far removed from the context at hand that
they are most likely to be wrong ones to use. Let’s illustrate this pitfall with a deceptively
simple problem.

Steam is generated in a batch boiler at the rate of m 2 = 20 kg/s. (See Figure 7.11.)
Problem: What is the heat required q1 ?

Table 7.2: Thermodynamic states and parameters for the process in Figure 7.11

Vapour Liquid Point 2


H, kJ/kg 2746.4 632.2 2746.4
U, kJ/kg 2559.5 631.7 2559.5
V, m3/kg 0.3928 0.00109 0.3928
Hold-up, kg (variable) (variable) ----
Flow, kg/s ---- ---- 20

Reference: H 0 (saturated water @ 15 °C) =63.0 kJ/kg


S 0 (saturated water @ 15 °C) = 0.2244 kJ/kg K

We look at two expressions of the first law, ∆U = Q − W and ∆H = Q − WS , both of


which are reduced and truncated expressions. Note that in this example Q = Q1 , the heat
input to the boiling water, since Q = Q1 − Q2 and Q2 = 0 (since no heat is rejected by the
system). As used here, the quantities Q’s and W’s are expressed as per mass of steam
leaving: Q = q m 2 ; Q1 = q1 m 2 ; WS = w s m 2 . The system is at constant temperature and
pressure and constant volume, to boot. No shaft work is transacted between the system
and the surroundings. Some texts could be misinterpreted as suggesting that the
∆H expression is “good” for applications where T , P are constant; and the ∆U expression
is “good” for constant T , V . So, to find , we may be tempted to consider either of the
following, both being legitimate, widely used and familiar expressions of the first law of
thermodynamics:

∆U = Q (with W = 0 ); or

∆H = Q (with WS = 0 )

The first expression Q = ∆U doesn’t foot the bill. No work transacted means
only WS = 0 , not W = 0 . Since W = WS + ∆PV and ∆PV ≠ 0 , so W ≠ 0 . The second

154
expression Q = ∆H is beguiling. It follows from m 2 H 2 − m 1 H 1 = q − w S (obtained from
the first law). With m 1 = 0 (no material entering system) and w S = 0 (no shaft work
transacted), we have then from the first law m 2 H 2 = q ; or H 2 = q m 2 = Q , where Q is
the heat transacted on the basis of a unit mass of steam leaving. Now, this has gone from
beguiling to downright absurd, because the expression H 2 = Q is nonsensical. H 2 is a
point value, whose numerical value depends on the arbitrary reference of the particular
data source used. But Q can’t be arbitrary! The heat required to raise the given quantity
of steam can’t possibly depend on the arbitrary reference used in a certain data source
from which H is obtained.

Actually, we shouldn’t have pursued the relation Q = ∆H once we found out as we did
that ∆U = Q doesn’t foot the bill, because the two expressions we started with,
∆U = Q − W and ∆H = Q − WS , are identical but in different guises. If it doesn’t work in
one guise, it can’t possibly work in another. However, it must be emphasized that both
expressions are correct, but only in their proper context.

For our deceptively simple problem we actually have to invoke the full expression of the
first law (7.1):

 u2   u2  d
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g  = q − w S − (mCU C )
 2   2  dt

Then, in the context of the problem at hand, we reduce and truncate it to (see equation (1)
in Appendix IX):
d
q1 = m 2 H 2 + (mCU C ) (7.18)
dt

From equation (6) in Appendix IX we have:

d
(mCU C ) = m 2 U L − U V VL VV = -12527 kW; and
dt VL VV − 1

m 2 H 2 = 54928 kW; this results in q1 = 42401 kW; or

Q = q1 m 2 = 2120.1 kJ per kg of steam leaving.

As an approximation and sanity check, we can say that, ignoring the effect of the varying
accumulation of liquid and vapour within the constant volume system, the rate of steam
production equals the rate of boiling. The latent heat of vaporization given by
H V − H L = 2114.2 kJ/kg is pretty close to Q =2120.1 kJ/kg, the small difference being
the effect of increasing vapour accumulation, with concomitant decreasing liquid
accumulation, inside the system. The quantity U V − U L = 1927.8 kJ/kg is not only the

155
wrong answer for Q , it is also meaningless in the context of our problem. The answer
Q = H 2 = 2746.4 kJ/kg (on the arbitrary reference basis of the one particular data source
used here) is beyond wrong and meaningless; it is disastrous! [ H 2 obtained from another
data source is different; for example HYSYS© steam data base gives H 2 = -13179.3
kJ/kg, simply because steam enthalpy has different reference zero there.]

When working with thermodynamics it is important not to use the truncated expressions
of the first law (H =Q – W S or U = Q – W) in an inappropriate context.

7.14 A little Historical Divertimento

Horse-powered calorimetry

James Prescott Joule (British physicist and brew master, 1818-1889) was a great
experimentalist. Among his great works were experiments he devised and carried out to
establish the work-heat equivalence, known as the “mechanical equivalence of heat”. One
of his experiments of this genre is depicted in Figure 7.12.

W S 1
A
Legend:

E C A: Drive shaft and spokes assembly

D B: Friction tube
B
C: Thermometer
Elevation view D: Insulated calorimeter walls
Horse-drawn E: Water level

Horse-drawn
Plan view
Figure 7.12:
Joule's horse-powered calorimetry (rendition with artistic license)

The work to drive the shaft was provided by a team of two horses trotting in a circle. The
shaft work is converted to heat by means of the friction tube. From the pre-measured
torque Γ of the shaft and the average rotational speed ω , the rate of shaft work production
is W S = Γω . The total shaft work done W S for the duration ∆t of the experiment is Γω∆t .
From the known mass m of water, its specific heat C P and the recorded temperature
rise ∆θ , the heat absorbed is mC P ∆θ .

156
It is said, this historic experiment was described in a lecture in a certain prestigious
English university. At the end of his description, the lecturer asked his audience, “What
do you think James Prescott Joules saw in all this?” The students replied in unison, “A lot
of horse manure!”

Apart from the obvious stated by the students, nothing else about this experiment is quite
that obvious. What Joule could conclude from it is that Γω∆t of work is equivalent to
mC P ∆θ of heat. But equivalence is not the same as equality. The difficulty was
aggravated by the fact that heat and work were expressed in different units: heat in Btu or
calorie and work in foot-pound-force, foot-poundal, horse power-hour, etc. If, however,
Γω∆t = mC P ∆θ , or work indeed equals heat, then the implication is profound. It means
that mechanical shaft work can be converted completely into heat. Joule took a leap of
faith and subscribed to the idea that work can be converted completely to heat. Many of
his collaborators and contemporary leapt with him. To Joule the difficulty presented by
disparate units of measurement, which prevented the immediate recognition of this
equality, was only superficial. This prompted him to champion a system of consistent
units, the precursor of our SI units, in which the hero is immortalized in the unit of
energy, the joule (J). It is hard to ascertain if Joule did in fact work with such a divinely
comical contraption as the horse-drawn calorimeter. He did, however, perform many
calorimetric measurements to establish the so-called work-heat equivalence (equivalence
is a misnomer; he was actually out to establish the equality), using various experimental
set-ups; some of these set-ups, for their great pedagogical values, are mimicked in high
school labs today (with instrument far more precise than what was available in Joule’s
days). But then, if you are a person of Joule’s caliber, there are bound to be some
anecdotes, plausible as they are comical, attributed to you, usually without your
knowledge or consent, because anecdotes of this sort gain currency only after your
demise.

Today, armed with hindsight and a system of consistent units like the SI, as well as with
results of great precision obtained at high school labs, we know Joule was right. Yes,
work can be converted completely to heat, joule for joule.

Cerebral heat engine

Like Joule, Nicolas Léonard Sadi Carnot (French physicist and engineer, 1796-1832)
thought a great deal about heat and work. Unlike Joule, Carnot posed his question in
reverse. Paraphrased and cast in the nomenclature used in this chapter, his question was:
Given a quantity of heat of Q1 supplied at temperature θ1 , can it be converted completely
into a quantity of shaft work WS ? His answer was: No, a certain amount of heat Q2 must
be rejected at temperature θ 2 to accompany the production of shaft work WS . This partial,
and it can only be partial, conversion is carried out in a theoretical construct known (to
us) as the Carnot heat engine, as shown in Figure 7.13.

157
Given heat Q1 supplied at θ1 , only part of it can be converted to shaft work WS and the
remainder Q2 must be rejected at θ 2 , with θ 2 < θ1 , so that: WS = Q1 − Q2

Source at θ1
Surroundings

Heat engine System


WS

Sink at θ2
Figure 7.13:
Carnot heat engine

Lopsided Mother Nature: heat-work asymmetry and entropy of heat

Work can be converted completely to heat; but heat can only be, at best, partially
converted to work. As a matter of fact, Q1 at θ1 can be completely rejected as Q2 at
θ 2 without producing any work! In that case, Q2 = Q1 and WS = 0 . This glaring
asymmetry of heat-work conversion got Rudolf Julius Emanuel Clausius (German
physicist and mathematician, 1822 – 1888) thinking hard. If you look at a picture of
Clausius (or some rendition of his likeness), you will know that there is a great deal
cooking behind those thick eyebrows, accentuated by that permanent frown on his face;
above those eyebrows protrude a broad forehead and a bulging cranium. When you get a
guy like that thinking, you know for sure something earth-shattering is going to come out
of it.

An equation like WS = Q1 − Q2 is a statement of energy balance. The reason no one can


ever devise an engine to achieve the complete conversion of heat to work is that,
according to Clausius, there is a qualitative and fundamental difference between heat and
work. He made one bold conjecture, hallmark of a genius, that given a quantity of heat
Q available at temperature θ (on the absolute scale), the ratio Q θ is a measure of its
quality; work WS is equal to heat available at infinitely high temperature.

Clausius went beyond the conjecture and set out to “explain” Carnot’s theorem. He
pictured work and heat as two forms of energy. Work is perfectly ordered, its disorder

158
being zero. Heat, on the other hand, is inherently disordered, to various degrees. The
measure of disorder is Q θ , the higher θ the more ordered is Q . In a process of energy
transformation, we start with a certain amount of energy imbued with a certain amount of
disorder; and we always end up with the same amount of energy but with an amount of
disorder greater than the one we started with. At best, only in an ideal process could we
preserve the original disorder. Clausius called this measure of disorder entropy, denoting
it with the symbol S . By this reasoning, we have the entropy S Q of a quantity of heat
Q as: S Q = Q θ . Applying this to the Carnot heat engine, we have, in line with Clausius’s
thesis:

WS = Q1 − Q2 (Energy conservation in heat-to-work conversion)


Q2 Q1
− ≥ 0 ; or, S Q 2 − S Q1 ≥ 0 (Entropy increase in heat-to-work conversion)
θ2 θ1

We recognize immediately that, stated on a unit time interval, the first relation is a
specialized expression of the first law and the second the specialized expression of the
second law. The first and second law are given in (7.1) and (7.5), which we repeat below
for ease of reference:

First law

 u2   u2 
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g 
 2   2 

d
( )
= q1 − q 2 − (w S 2 − w S 1 ) −
dt
(mCU C ) (7.1)

Second law
d
(mC S C ) ≥ q1 − q 2
 
(m 2 S 2 − m 1 S1 ) + (7.5)
dt θ1 θ 2

For the Carnot heat engine, m 1 = m 2 = 0 (no material entering or leaving);


d (mCU C ) dt = d (mC S C ) dt = 0 ( mC = 0, no material held up inside system). On a unit
time basis, we have according to the nomenclature and convention adopted for this
chapter:

q1 → Q1 ; q 2 → Q2

w S = w S 2 − w S 1 → W S

For an ideal engine, the equality in the second law expression holds and the work
produced WS (ideal ) is then the maximum possible:

159
θ2
Q2 = Q1 (ideal engine)
θ1
θ −θ2
WS (ideal ) = Q1 1
θ1

Note the subtlety that even for an ideal engine that yields the maximum possible shaft
work W S (ideal) a quantity of heat Q 2 must still be rejected. Maximum possible shaft
work is not obtained by converting all of Q 1 to work, but by incurring no entropy
increase, i.e. by having Q2 θ 2 − Q1 θ1 = 0. In other words, Carnot’s theorem (Q 2 > 0) is
true whether the engine is ideal or real.

It is easy to see that Carnot’s theorem that a finite non-zero quantity of heat Q2 must be
rejected at temperature θ 2 is a loud statement of the second law as we now know it,
whether or not Carnot actually knew it. By Clausius’ analysis, we have the first and
second law applied to the Carnot engine as: WS = Q1 − Q2 and Q2 θ 2 − Q1 θ1 ≥ 0 . Note
that the terms Q2 θ 2 and Q1 θ1 are positive definite. If, contrary to Carnot’s theorem, all
of Q 1 were converted to work, we would have WS = Q1 and Q2 = 0, and we would be left
with − Q1 θ1 ≥ 0 . This is absurd, because a negative number (or rather the negative of a
positive definite number) cannot possibly be greater than or equal to zero! What is more,
it would violate the second law, which requires Q2 θ 2 − Q1 θ1 ≥ 0 ; this requirement
cannot be met unless Q 2 > 0.

Entropy of material substance

Clausius’ thesis: A given quantity of heat Q has entropy S Q = Q θ . This got Ludwig
Eduard Boltzmann (Austrian physicist, 1844 – 1906) thinking about the entropy of a
material substance. Question: Given a material with energyU , how much of this energy
is convertible to work? That depends on the other state function of the material, namely
entropy S . A material can have the same energy U at two (or more) distinct states: say,
state 1 (T1 , P1 ) and state 2 (T2 , P2 ) . That is, we have U (T1 , P1 ) = U (T2, , P2 ) . We have then
perforce S (T1 , P1 ) ≠ S (T2 , P2 ) , because the two states are distinct. Answer: U in that state
which has the lowest entropy S has the most capacity for conversion to work. The answer
is predicated on S (T1 , P1 ) ≠ S (T2 , P2 ) . What if S (T1 , P1 ) = S (T2 , P2 ) ? If so, then, the states
(T1 , P1 ) and (T2 , P2 ) are identical, i.e. T1 = T2 and P1 = P2 , not distinct as stipulated in our
premise. Recall that the state of a material is uniquely determined by the two state
functions (U , S ) .

Ludwig Boltzmann is your quintessential tragic figure who belongs in an opera. He


worked out the fundamental expression for entropy of a material substance as S = k ln W ,
based on his own kinetic theory. (W is the number of distinguishable ways in which the

160
particles in an ensemble can be arranged (or distributed) subject to a given total energy of
the ensemble.) This supremely simple and elegant expression turns out to be correct. The
problem is that the kinetic theory hinges on the physical reality of atoms (and molecules).
In his days, the reality of atoms and molecules as constituents of material substances was
a touchy issue. Among his detractors was a figure no lesser than Ernst Mach (of the Mach
number fame), who denied the reality of the atom (hear this!) on philosophical ground.
Shakespeare has a name for an authority of this category, as he puts it brilliantly in King
Lear:

Court Jester: The reason why the seuen Starres are no mo than seuen, is a pretty one.
King, wagging sagely: Because they are not eight.
Court Jester: Yes indeed, thou wouldst make a good Foole.

Boltzmann incurred the wrath of the Viennese scientific establishment. He was allowed
to use the word atom provided he dismissed it beforehand as a mere mathematical
construct in his equations and not treated it as something physically real, all because the
sage Mach decided that atoms did not exist because atoms did not exist. Compounded by
his bipolar disorder, Boltzmann was driven to severe depression and took his own life.
The reality of atoms and molecules as the constituents of material was put on firm ground
by Albert Einstein in his analysis of the Brownian motion. This analysis was published in
1905; but Boltzmann didn’t live long enough to read it. You see, mixing philosophy with
science is as tragic as mixing religion with politics. As Canadian comedian Russell Peters
intones it in his lyrical East Indian accent, “Somebody is going to get hurt!”

161
Chapter 8

REFRIGERATION

8.1 Introduction

In this chapter we present the mechanical refrigeration and the absorption refrigeration.
We show how thermodynamics applies to these processes in all its simplicity,
transparency and elegance. We discuss also some of the common features that exist in the
commercial designs of the mechanical refrigeration – features that greatly impede the
efficient operations of a refrigeration unit.

8.2 Mechanical Refrigeration

The mechanical refrigeration (also known as vapour compression refrigeration) is shown


in the conceptual scheme in Figure 8.1.

Surroundings
Heat rejected q 2
(Condenser duty)

System Pressure difference PH − PL is


Vapour
Condenser @ TH maintained by the compressor
Pressure PH Liquid Refrigerant:
w S 1 boiling in evaporator at TL , PL ;
condensing in condenser
Shaft w ork Pressure PL at TH , PH
Com pressor

Vapour : Letdow n valve


Evaporator @ TL (isenthalpic throttling)

First law: w
 S 1 = q 2 − q1 q1 Heat absorbed
(Refrigeration duty)

Figure 8.1:
Schematic of mechanical refrigeration (a.k.a. vapour compression refrigeration)

This is a common method of achieving refrigeration at industrial scale and in the


household refrigerator and air-conditioner. The process is relatively easy to understand.
Heat is removed at a low temperature from the object of refrigeration (from the air space
and contents in a refrigerator or from a process stream) by boiling at temperature TL a
refrigerant in a heat exchanger generically called the evaporator. The evaporator will be
the chiller tubes in gas processing, or the coils in the freezer compartment of the
household refrigerator. The refrigerant vapour arising from boiling is compressed and
condensed by removing heat from it at temperature TH in another heat exchanger called
the condenser. In an industrial installation, the condenser is quite often a forced air
cooler. In the household refrigerator, the condenser manifests itself as the coils mounted

162
at the back of the refrigerator, where heat rejection is effected by natural convection into
the room air.

First law analysis of mechanical refrigeration

From the full expression of the first law of thermodynamics (7.1):

 u2   u2  d
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g  = (q1 − q 2 ) − (w S 2 − w S 1 ) − (mC U C )
 2   2  dt

The derivation of the expression is given in Chapter 7. For convenience the symbols and
nomenclature are repeated here:

m : mass flow rate of material


P: pressure
q rate of heat transaction
U: internal energy per unit mass of material (refrigerant)
u: flow velocity of material
V; volume per unit mass
w S : rate of shaft work transaction
z: elevation of a point in the material above a common arbitrary datum level in the
gravitational field

Subscripts

1: entering the system, or from surroundings to system in the case of heat or work
transaction
2: leaving the system, or from system to surroundings in the case of heat or work
transaction
C: of material enclosed (or held up) in the system

For a closed system, such as the mechanical refrigeration, it is recognized that:


m 1 = m 2 = 0 (no material entering or leaving the system);
d (mCU C ) dt = U C dmC dt + mC dU C dt = 0 , since dmC dt = 0 (constant refrigerant
hold-up in the system), and dU C dt = 0 (constant state of refrigerant at any point on its
circulating path); wS 2 = 0 (no shaft work done by the system).

For the mechanical refrigeration the first law expression is reduced to:

w S 1 = q 2 − q1 (8.1)

Note that the reduced expression (8.1) is still exact, no terms having been deleted as
insignificant; all terms eliminated from the full expression are truly zero.

163
Second law analysis of mechanical refrigeration

From the full expression of the second law of thermodynamics (7.5):

d
(mC S C ) ≥ q1 − q 2
 
(m 2 S 2 − m 1 S1 ) +
dt θ1 θ 2
The derivation of this second law expression is given in Chapter 7. The meaning of the
symbols and nomenclature are repeated above under First law analysis of mechanical
refrigeration. The new symbols are:

S: entropy per unit mass of material


θ1 : absolute temperature at which heat is absorbed from the refrigerated load by the
system (i.e. source temperature), or temperature to which a load is refrigerated
θ2 : absolute temperature at which heat is rejected into the surroundings by the system
(i.e. sink temperature), or temperature of the surroundings

For the closed system in question, it is recognized that: m 1 = m 2 = 0 (no material entering
or leaving the system); d (mC S C ) dt = S C dmC dt + mC dS C dt = 0 , since
dmC dt = 0 (constant refrigerant hold-up in system), and dS C dt = 0 (constant state of
refrigerant at any point on the circulating path).

For the mechanical refrigeration the second law is reduced to:


q 2 q1
≥ (8.2)
θ2 θ1

Note that this reduced expression of the second law is still exact, no approximation
having been made by ignoring any terms considered insignificant in the full expression.
Note also that θ1 , θ 2 are the source and sink absolute temperatures of q1 , q 2 , respectively.
In the context of refrigeration, θ1 is the temperature at which the refrigeration load
transacts heat with the boiling refrigerant. For a household refrigerator, θ1 is the
temperature of its inside space (where food and drinks are kept). For the raw gas chiller
in gas processing θ1 is the temperature of the raw gas side; θ 2 is the temperature at which
the heat from the condensing refrigerant vapour is rejected to the surroundings, i.e. it is
temperature of the surroundings.

Why bother with the second law?

The first law doesn’t tell us how efficient our refrigeration process is; the second does.
The equality of (8.2) holds for an ideal process, one that functions at 100 % efficiency.
For a given refrigeration load of q1 , the condenser duty q 2 and the shaft work w S 1 for the

164
compressor would be minimum if the refrigeration process were ideal. Let’s denote these
minima as q 2 (min ) , w S 1 (min ) . From (8.2) we have:
θ2
q 2 (min ) = q1
θ1

This, together with (8.1), gives:

θ 2 − θ1
w S 1 (min ) = q1
θ1

The actual shaft work for the compressor is of course greater than the minimum, i.e.
w S 1 > w S 1 (min ) , because our process is real, not ideal, due to factors that include barriers
to heat transfer and friction losses in the compressor. From this we can state the
efficiency of the process any way we like. For example, if the ratio w S 1 w S 1 (min ) = 1.38,
we say that our process uses 1.38 times more shaft power than the ideal; and it also
indicates the margin of improvement to which we can aspire. Or, equivalently, the ratio
w S 1 (min) / w S 1 = 0.72 says that we would use only 72 % of the actual shaft power for the
same refrigeration load q1 if our refrigeration machine were ideal (which it is not)!

Note that the first law expressed in (8.1) is true for all processes, reversible (ideal) or not.
Substituting the ideal compressor shaft power and ideal condenser duty into (8.1), we
have w S 1 (min) = q 2 (min) − q1 . Subtracting this from (8.1), we have
w S 1 − w S 1 (min) = q 2 − q 2 (min) > 0 . This means that the condenser duty is higher for a
real process than for the ideal; or, q 2 > q 2 (min) as we already know intuitively.

It is curious to note that many texts on refrigeration adopt a long-winded way to state the
efficiency of refrigeration. There the “concept” of the coefficient of performance (COP)
is introduced; it is defined as:

q1
COP =
w S 1

From this it can be seen that coefficient of performance of the ideal is:

q1 θ1
COP ideal = =
w S 1 (min) θ 2 − θ1

It is seen that COP ideal > COP. The ratio of these two COPs is then used to indicate the
efficiency of the real machine by, in effect, comparing w S 1 (min) and w S 1 ; as we can see,
COP/COP ideal = w S 1 (min) / w S 1 . It would be a lot more straight forward and instructive

165
just to compare w S 1 and w S 1 (min) directly without going through the superfluous and
long-winded “concept” of COP.

8.3 Absorption Refrigeration

We can see that the refrigeration process is really the Carnot engine (discussed in Section
7.14) operating in reverse, as shown in Figure 8.2.

Sink at θ2

Heat rejected q
2
(condenser duty)

Shaft w ork
Reverse Carnot engine:
w S 1 Refrigeration or heat pump System

Heat absorbed q1


(refrigeration duty) Surroundings

Source at θ1 First law: w


 S 1 = q 2 − q1

Figure 8.2:
Carnot heat engine in reverse: principle of refrigeration (or heat pump)

Shaft work w S 1 is done to the system to remove heat q1 (refrigeration duty) at a lower
temperature θ1 and to reject heat to the surroundings at the rate of q 2 (condenser duty) at a
higher temperature θ 2 , such that q 2 = q1 + w S 1 by dint of the first law. If we couple the
“normal” Carnot engine with one operating in reverse, we can get rid of any transaction
of shaft work between the process (system) and the surroundings. This is the basis and
inspiration for the absorption refrigeration process, as shown in Figure 8.3, in which no
shaft work is transacted between the system and the surroundings.

In Figure 8.3 it is indicated that the shaft work output w S 1 of the normal Carnot (heat
engine) becomes the shaft work input to drive the reverse Carnot (refrigeration). It should
be noted that this internal exchange of work is strictly conceptual; i.e. conceptually, work
is produced and consumed entirely within the system without any transaction between the
system and the surroundings. In the physical realization of the absorption refrigeration,
there is no need to actually produce and consume shaft work internally, as will be seen
shortly.

Thermodynamics only governs the transactions of materials, heat and work between the
system and the surroundings; it has no bearings on the details and contrivances that
physically constitute the system. In other words, thermodynamics tells us what can or

166
cannot be done; it doesn’t tell us how it can be done. How it can be done is left to our
own devices and ingenuity. The conceptual scheme in Figure 8.3 says we can devise a
gizmo to refrigerate at the temperature of θ1 with only heat input but no shaft work input
to the system; in short, to achieve refrigeration, all we need is a heat supply q A1 available
at a certain temperature θ A1 ! This means that we can use low level waste heat to achieve
refrigeration. (The temperature θ A1 turns out to be quite low, 80 to 150 °C given the
common commercial refrigerants available.) This is an amazing and irresistible concept
that has inspired some of our most brilliant minds: Albert Einstein and Leo Szilard
actually took out a joint patent on a scheme for pumpless absorption refrigeration, in
which they claim (legitimately, of course) to achieve refrigeration with nothing but a
source of heat! (See Einstein, Szilard (10).)

Source Normal Carnot Reverse Carnot Sink


(heat engine) (refrigeration)
at θ Α 1 at θ 2

Heat rejected q
2 Surroundings
q A1
(Condenser duty)

Carnot w S 1 Reverse Carnot Engine: System


Engine Refrigeration or heat pump
Shaft w ork

Heat absorbed q1


q A2
(Refrigeration duty)
Sink Source
at θ Α 2 at θ 1

First law: q1 + q A1 = q 2 + q A 2


Figure 8.3:
Conceptual scheme of absorption refrigeration by coupling of a Carnot engine and a
reverse Carnot engine. No shaft work is transacted between system and surroundings.

In most absorption refrigeration schemes at industrial scale, as shown in Figure 8.4,


however, a pump is required to circulate the refrigerant solution because the condensation
of the refrigerant has to occur at pressure P H and its evaporation at P L , with P H higher
than P L , just as it is the case in the mechanical refrigeration process. An ammonia-water
solution is commonly used in the absorption refrigeration; and we will use this refrigerant
in our illustration.

Note that as marked in Figure 8.4, the configuration of equipment to the left of vertical
broken line - i.e. regenerator, absorber, cooler, heat recovery exchanger, pump and,
distinctively, heat input q A1 to the regenerator - takes the place of the compressor in the
mechanical refrigeration shown in Figure 8.1. It is, in fact, exactly equivalent to the
compressor. The heat supply q A1 is used to drive out ammonia vapour from the rich
ammonia solution entering the regenerator. The configuration of equipment to the right of
this line is identical in the two schemes: mechanical and absorption. What we achieve
with the scheme in Figure 8.4 is to replace the compressor and its large power

167
requirement with a source of (waste) heat q A1 plus a small pump moving liquid, hence
with a small power requirement.

Equivalent to Heat rejected q


2 Sink

Surroundings
compressor (Condenser duty) at θ 2

Vapour
System NH 3 Condenser
Source q A1 Regenerator
at θ Α1 (heat supply)
Liquid
Lean NH 3 Pressure PH NH 3
Heat solution
Sink q A2 Recovery
Cooler
at θ A2
Rich NH 3
Vapour
Pressure PL
w S 1 solution
NH 3 Evaporator
Pum p Absorber

Heat absorbed q1 Source


First law: q1 + q A1 + w S 1 = q 2 + q A 2 (Refrigeration duty) at θ 1
Isenthalpic throttling
Figure 8.4:
Schematic of industrial absorption refrigeration with pump

The use of a pump requiring external shaft work is the fly in the ointment, even though
the shaft work required for the pump is only a small fraction of the shaft work required to
drive the compressor in an equivalent mechanical refrigeration system. We know that we
can do without the pump because we are assured by the conceptual scheme shown in
Figure 8.3, which says no input of shaft work is needed. To dispense with the pump we
just have to concoct a device in which the refrigerant (NH 3 in our example) can condense
in the condenser at the same pressure as that at which it boils in the evaporator; in other
words, we have to let the evaporator and the condenser operate at the same total pressure,
bearing in mind that the condenser operates a higher temperature (50 °C, for example) to
reject heat to the surroundings and the evaporator has to operate at a much lower
temperature (-10 °C, for example) to achieve the desired refrigeration. As it turns out, an
arrangement, as shown in Figure 8.5, to accomplish this is rather ingenious: it is the
epitome of simplicity and elegance.

The scheme here (Figure 8.5) is to spike the ammonia-water system with an inert gas,
such as hydrogen. An inert gas in this context is a gas that has a low solubility in water
(compared to that of ammonia in water), remains gaseous and does not react chemically
with ammonia or water under the operating conditions of the system. (Helium is another
good inert for this purpose.) In the vapour space between the regenerator-cum-separator
and the condenser only ammonia vapour is present; whereas the vapour space between
the evaporator and the absorber is occupied by a mixture ammonia vapour and hydrogen
gas (inert). Both vapour spaces have the same total pressure. For an ammonia-water
system, this total pressure may be 2000 kPa (a). Without the inert present there, the total
pressure is also the partial pressure of ammonia vapour in the condenser; this allows the

168
ammonia vapour to condense at 50 °C at this total pressure (which is also the partial
pressure of NH 3 in the condenser), making it feasible to use an air-cooled condenser. The
closed system is doped with the right quantity of inert so that the partial pressure of
ammonia in the vapour space between the evaporator and the absorber is maintained at
300 kPa (a), for example, at the prevailing total pressure of 2000 kPa (a). The inert
present in that vapour space is such that the mole fraction of NH 3 there is about 0.15 (=
300/2000).This partial pressure allows the liquid ammonia in the evaporator to boil at -10
°C, given that the boiling point of NH 3 is -10 °C at 300 kPa (a). This ingenious scheme is
founded on the understanding of phase equilibrium, whereby phase change
(vaporization/condensation) is driven by partial pressure, not total pressure. (Strictly
speaking, we should say driven by fugacity, not total pressure. Fugacity is the effective
partial pressure, effective because it accounts for the effects of intermolecular forces for a
real gas.)

Sink
q 2 (condenser duty)
Vapour NH3
at θ 2 Draft Tube/
Small continuous Condenser
vapour purge Heating Surface
NH3 vapour space
Regenerator
Source H2+NH3 vapour space
U liquid /Separator
at θ 1 Vapour NH3
seal
Evaporator Cooler
Liquid NH3
Lean NH3
q1 (refrigeration duty) solution

Absorber

Rich NH3 solution


Sink
Heat Recovery
at θ A2
q A2 (heat removed by cooler)
Flame
Source q A1 (heat absorbed from flame)
at θ Α1
First law: q1 + q A1 = q 2 + q A 2
Figure 8.5:
Pumpless absorption refrigeration; only heat, no shaft work, is transacted
between the process and the surroundings

As points of design details (see Figure 8.5), the small continuous vapour purge from the
condenser vapour space to the absorber vapour space is to prevent the accumulation of
inert in the vapour space between the regenerator/separator and the condenser, since such
accumulation would reduce the partial pressure of ammonia in this vapour space and
impede its condensation. The inert (hydrogen gas) has a very low but not zero solubility
in water. Entrained in the circulating solution, the inert would be driven off by the heat
applied to the regenerator, get trapped and grow in quantity in the condenser vapour
space, thereby reducing the partial pressure of ammonia in the condenser and hindering

169
its condensation. The heat recovery exchanger is just an economizer to reduce
q A1 and q A 2 ; it has otherwise no bearing on the principle of the process. It should be noted
that the circulation of ammonia-water solution (between the absorber and regenerator) is
sustained by the thermosiphon effect created in the regenerator by the external heat
source marked “Flame” in Figure 8.5; also, since the whole closed system operate at one
single total pressure, there is no need for any pressure letdown valve anywhere in the
circuit. The two operating vapour spaces, NH 3 and NH 3 + H 2 , are kept segregated by the
U-liquid seal through which the liquid ammonia drains from the condenser into the
evaporator; and by the liquid leg of rich ammonia solution draining from the absorber to
the regenerator.

The pumpless absorption refrigeration process is used in portable or domestic


refrigerators in locations where no electrical power is available as motive power for the
compressor required in the mechanical refrigeration process - in remote corners of the
world or in a modern mobile camper. In large commercial or industrial installations,
absorption refrigeration is used where large quantity of low-temperature heat (or waste
heat) is available. In these installations, a circulation pump is used, as shown in Figure
8.4. The disadvantage of having a small pump is far outweighed by the delicate hydraulic
balance required of a pumpless arrangement. In these large scale installations, more often
than not multiple evaporators at various elevations and situated at locations over a large
plot area are required to perform separate refrigeration duties. Maintaining circulation of
the refrigerant by hydraulic balance without a pump would be impractical, if not
impossible, in such applications.

8.4 Cascade Mechanical Refrigeration

q 2 (condenser duty)

Condenser
w S 1 A
-40 oC Propane loop 43 oC
110 kPa a 1471 kPa a

w S 1B
-85 oC Ethane loop -30 oC
120 kPa a 1000 kPa a
Evaporator
q1 (refrigeration duty)
: Letdow n valve
First law: q 2 = q1 + (w  S 1B )
 S1 A + w (isenthalpic throttling)

Figure 8.6
Cascade mechanical refrigeration:propane-ethane cascade;
refrigeration temperature - 85 oC

170
With two mechanical refrigeration cycles cascaded, deep cryogenic temperatures can be
achieved. A propane-ethane cascade (Figure 8.6), for example, can achieve the cryogenic
temperature of -85 °C in the evaporator.

The evaporator of the propane loop acts as the condenser for the ethane loop.

8.5 Refrigeration in Gas Processing

In gas processing refrigeration is used to remove heavier components (some ethane,


propane and heavier) from dehydrated raw gas. The removal of these components from
the raw gas reduces its hydrocarbon dew point making it suitable for transmission by
pipeline. In addition there are installations known as “straddle plants” which intercept
pipeline gas (which already meets pipeline hydrocarbon dew point) to remove the
remnant ethane and propane it contains. The removal of these remnants is accomplished
by cryogenic cooling using cascade mechanical refrigeration and turbo-expansion,
usually in combination. Turbo-expansion is a process in which deep cooling of a gas
initially at high pressure is achieved by letting down the gas pressure through a turbine.
This deep cooling of the gas is effected mainly by the shaft work output of the turbine
and, to a lesser extent, by the Joule-Thompson effect of pressure letdown. (The turbo-
expanded gas, with the ethane and propane removed, needs to be recompressed to
continue its pipeline transmission; part of the power for this recompression is supplied by
the turbo-expander.) This “last squeeze of the juice” out of the pipeline gas is justified
by the market values of ethane and propane as petrochemical feedstocks.

8.5.1 Design Parameters for Mechanical Refrigeration

We shall present the typical mechanical refrigeration for raw gas processing (reduction of
hydrocarbon dew point). It is a simple process. The main thrust of the presentation here is
to demonstrate how a simple process can be designed and set up to fail when the
underlying principles and phenomena governing its operations are misunderstood or
ignored.

Coolant for condenser

For our illustration here, we use ambient air as the coolant for the condenser. This is
typical of many gas processing facilities where a central cooling water system is not
justifiable. The ambient air temperature then dictates the operating pressure achievable in
the condenser, which in turn determines the discharge pressure of the compressor. In
locations, such as in Alberta, the variations of ambient air temperature are extreme: from
33 °C to -45 °C. For the purpose of the design and sizing, a judicious choice of the
ambient air temperature as well as the condensing temperature has to be made in the
context of this large variation in actual ambient temperature.

171
Refrigerant

The common refrigerant propane is used in this illustration. The choice of refrigerant
determines the temperature at which refrigeration can occur with an adequate positive
pressure maintained at the suction of the compressor. This compressor suction pressure is
usually maintained at about 10 kPa above the atmospheric pressure to prevent the ingress
of air (through leaks) into a closed system of refrigerant. The lowest pressure in the
closed mechanical refrigeration system exists at the suction to the compressor.

From these design considerations, we can establish the following process parameters for
the sizing, design and specifications of equipment:

Refrigerant: propane
Ambient air temperature: 29 °C (Alberta, Canada)
Condensing temperature: 43 °C (propane condensing temperature)
Compressor suction pressure: 110 kPa (a)
Compressor discharge pressure: 1471 kPa (a) (bubble point pressure at 43 °C)
Chilling temperature: -40 °C (bubble point temperature of propane at
110 kPa (a))

The choice of 29 °C for ambient air temperature and 43 °C for condensing temperature is
considered optimal based on industry experience in Alberta. Extreme temperatures
higher than 29 °C do occur, but rarely. If, for example, 33 °C (instead of 29°C) were used
as design basis, the air cooler tube area would be about 40 % greater than that based on
29 °C. At the few locations where such extremes do occur, they last just a few hours a
year on the average. Over-sizing by such a large margin to cater for those rare few hours
is not justified. The choice of 43 °C is for condensing temperature gives a temperature
approach of 14 °C for sizing the air cooler. The use of a lower condensing temperature of
39 °C, for example, would increase the tube areas of the air cooler by about 40 %. The
savings in compression power by 5% due to the lower condensing pressure of 1400 kPa
(a) (corresponding to 39 °C) hardly justifies such an over-size. At any rate, in Alberta,
more than 95 % of the time we can achieve an operating condensing temperature below
the 43 °C; more than 50 % of the time we have to shed the air cooler fan power to prevent
the condensing temperature (hence condensing pressure) from falling too low. As will be
seen in the Section 8.5.3 on process control, a minimum pressure has to be maintained in
the condenser and refrigerant accumulator to make it possible to deliver the refrigerant to
the users.

8.5.2 Equipment Design

The process scheme is shown in Figure 8.7.

172
Condenser

It must be recognized that total condensation occurs in the condenser. This means the
vapour flow at the outlet of the condenser is zero. Unlike partial condensation, there is no
vapour flow to sweep the liquid out of the condenser tubes. The condenser should then
have a single-pass tube bundle with a slope (5 to 10 degrees) down toward the outlet to
allow drainage of the liquid out of the tubes. In a horizontal tube, a layer of liquid would
have to build up in the tube to provide the hydraulic gradient in order to drive the
drainage. For a 8 m long horizontal tube, the liquid hold-up (or flooding) in the tube
required to establish the hydraulic gradient can easily cover 30 % or more of the tube
area. Flooded tube surface is not accessible to the condensing vapour, resulting in the loss
of effective heat transfer area.

PD com pressor Desuperheat spray


(Screw type) (narrow cone nozzle)
Condenser
Oil
Separator

Oil Cooler Equalizing


Plenum
M M line
Liquid
PC2
Raw gas KO Drum
FI ** PC1

LC

M
Chiller Accum ulator
Gas+liquid
** PC1 logic includes fan speed sw itching Spray Pum p
strategy and louver manipualtion
Figure 8.7:
Gas processing: reduction of hydrocarbon dewpoint in raw gas

Pre-saturation spray

The size (or face area) of the condenser tube bundle and fan coverage can be reduced
considerably if the vapour enters the condenser saturated, so that no desuperheating is
required in the condenser. The compressor discharge may be superheated by as much as
20 °C. To desuperheat this in the condenser would require an inordinately large extra
tube area. It is highly inefficient to desuperheat vapour in the condenser, or in any
exchanger intended for condensation. The superheated vapour from the compressor
discharge can be brought to saturation by spraying liquid propane into the flowing
vapour. See Figure 8.7. The equipment required for the spray consists of a small low head
pump (usually small enough to be mounted in-line), small piping (usually no larger than
2 inch) and a narrow-cone spray nozzle. The extra equipment is a very small prize to pay
for the savings in condenser size, fan coverage and fan power. The control of the pre-
saturation spray is simple; see Section 8.5.3 on process control.

173
Sizing of an air-cooled finned tube bundle

The face area of a finned tube bundle is the projected area of the bundle perpendicular to
the air flow (see Figure 8.8). This face area determines the fan coverage required for a
given application. For air flow sustained by fans, as is the case of most forced air coolers,
the optimal number of tube rows for a tube bundle is four. For a given tube area required,
two tube rows instead of four, for example, would double the face area and hence double
fan coverage; while an arrangement with more than four rows of tubes would reduce the
face area and fan coverage accordingly. However, for the number of tube rows greater
than four, the developed head required for the forced air flow would be so high that it lies
beyond the capability of a fan. In other words, with more than four rows of tubes a fan
simply cannot develop such a high head without overwhelming air slippage at the blades.
A fan is a simple machine that employs only rotor blades in a single stage to impart head
to the air; the imparted head manifests as fan discharge pressure. This simple construction
limits the head it can develop to less than 25 mm of water column. (Other mechanical
devices to develop heads to air and gases are blowers and compressors; these machines
accommodate multi-stage rotors, stator vanes and diffusers in their construction so that
they can develop much greater head than fans; but they are not economical in forced air
coolers.)

Tube bunddle
(optimal 4 tube row s)

A A

View A-A
Plenum
Face area = a x b
(space) Fan (typ.)
(or fan coverage area)
Fgure 8.8:
Face area (or fan coverage); with four tube rows (optimal), face
area, hence fan coverage, is proportional to tube area required

Refrigerant compressor

A positive displacement (PD) compressor is selected for this application. Unlike the
centrifugal or axial compressor, the PD compressor can ride on a wide range of discharge
pressure at a constant drive shaft speed. (By contrast, a centrifugal or axial compressor
can only operating against a narrow range of discharge pressure. These operating

174
characteristics are inherent and fundamental in the compressor type: PD, centrifugal and
axial.) With ambient air as the coolant for the condenser, the condensing temperature,
hence pressure, is expected to vary over a wide range, depending on the ambient
temperature, as discussed below. The ability of the PD compressor to ride on a lower
pressure obtainable in the condenser at a lower ambient temperature results in lower
discharge pressure, hence lower power consumption and a reduction of wear and tear of
the machine. By contrast, the use of a centrifugal compressor (or an axial compressor) for
this service would require the condenser pressure, against which it discharge, to be held
more or less constant.

8.5.3 Process Control

Compressor discharge pressure

The temperature at which condensation occurs in the condenser varies with the range of
ambient air temperature. This in turns means the prevailing pressure (bubble point
pressure of the refrigerant at the condensing temperature) in the condenser also varies.
This pressure can vary widely, for example: from 350 kPa (a) (corresponding to propane
condensing at -10 °C on a cold day) to 1471 kPa (a) (corresponding to propane
condensing at 43 °C on a hot day). The compressor discharge is allowed to ride freely on
this widely varying condenser pressure to take advantage of it. At the condenser pressure
of 350 kPa (a) the compressor power required is less than 20 % of that required at the
condenser pressure of 1471 kPa (a); this represents great savings in power and wear and
tear on the machine, provided of course that the compressor is allowed, by deliberate
design, to ride freely on this low pressure, and that PD compressor is used for this
service. Here we should gracefully accept a helping hand from Mother Nature, which she
extends on occasion by means of a low ambient temperature.

The pressure obtained in the condenser is also the prevailing pressure in the refrigerant
accumulator, the pressure being equalized in the vapour spaces of these two pieces of
equipment to facilitate draining of condensate from the condenser. (See Figure 8.7.) To
minimize compression power we would like to maintain the lowest possible pressure in
the condenser. The question on pressure control for the condenser boils down to: is there
a minimum operating pressure that must be maintained in the condenser? If so, what
process considerations determine it? The minimum operating pressure in the condenser
is one that is still adequate to deliver liquid refrigerant from the accumulator to the users,
as well as adequate to maintain compressor oil circulation when a screw type PD
compressor is used. Compressor oil circulation is elaborated in Section 8.6.2. This
minimum pressure depends on the locations and elevations of the refrigerant users
(chillers) and the required pressure difference (between compressor suction and
discharge) to maintain compressor oil circulation. The control strategy is to let the
pressure be whatever is achievable at a given ambient temperature, as long as it does not
fall below this minimum. One sensible way to implement this strategy is to shed the fan
power in response to falling pressure in the condenser (or accumulator). Many air-cooled
condensers use a bank of several fans, each fan equipped with the capability to make

175
multiple stepwise change in speed (0, ½, full speed, for example). Fan power shedding
can be executed by reducing the fan speed, stepwise, one fan at a time, until a step change
results in a stable condenser pressure above the minimum. In response to rising pressure
(due to ambient temperature rising) the fan speed change is executed in the reverse order,
namely stepwise increase of fan speeds. If a bank of 8 fans is used, each with a three
speed capability (0, ½ speed and full speed), we can make 24 stepwise changes in air
flow. The control target here is not a fixed pressure but one that is allowed to vary to
benefit from low ambient temperatures, but is maintained above the required minimum.
The control will of course work with variable speed fans, but given the control objective
here, which is easily achievable with stepwise adjustment of fan speed, the cost of the
hardware for variable speed motors is not justified.

In addition to fan speed, we also have the louver position (opening) that we can
manipulate. Note that the louvers should be in the fully open position when any of the
fans is blowing. (It does not make sense if the louvers were partially open, i.e. restricting
air flow, with any of the fans running.) At an extremely low ambient temperature coupled
with low refrigeration load (low plant throughput), it may occur that even with all the fan
speeds reduced to zero the condenser pressure still continues to fall due to the convective
air flow across the tube bundle, and imminently fall below the minimum allowable.
When this occurs, the air flow is throttled by reducing the opening of the normally fully
open louvers. Automation of this control strategy can be easily implemented in a control
scheme that coordinates the fan shedding and louver position manipulation.

Compressor suction pressure

The compressor suction is to be maintained at a pressure slightly above the atmospheric


to prevent the existence of vacuum (sub-atmospheric pressure) in the low pressure part of
the system, i.e. from the chiller to the compressor suction. Maintaining the internal
pressure above atmospheric prevents the ingress of air into the closed system; the ingress
of air can have dire consequences. This suction pressure control is accomplished with a
constant speed compressor equipped with a suction loader, or a variable vane to throttle
the flow to the compressor inlet.

Chiller

Chilling of the raw gas is effected by boiling the refrigerant in the chiller. The refrigerant
level in the chiller is maintained fixed so that the tube bundle remains totally submerged
in the boiling refrigerant. See Figure 8.7 for the level control for the chiller. The level
control valve lets the refrigerant into the chiller at a rate depending on the refrigeration
load, so that the chiller tubes remain submerged at all time. Only submerged tube surface
is accessible to the boiling liquid; tubes not submerged represent lost heat transfer area.

176
Pre-saturation spray flow

There is no need for automatic flow or pressure control of the pre-saturation spray.
Suffice it to have a flow indication and manual valve for flow adjustment. The flow is
adjusted to maintain the spray rate required in the most demanding case for complete
desuperheating (i.e. desuperheating to saturation) of the compressor discharge. In a less
demanding operating condition, this means that some liquid in excess is sprayed into the
compressor discharge flow and thus given a free ride through the condenser. For a well
drained condenser this excess liquid has no significant effect on the condenser operation,
because the total spray flow is small compared to refrigerant circulation rate.

Note that this ex situ desuperheating to saturation does not in any way change the
condenser duty q 2 , as determined by the first law expressed in (8.1). It only eliminates the
need to desuperheat using the condenser tube area ill-suited for the purpose. With this ex
situ desuperheating to saturation, the superheat contained in the compressor discharge is
in effect removed in the condenser by the mechanism of condensation, which is more
efficient, by at least an order of magnitude, than by removing it as sensible heat by
cooling the superheated gas phase to saturation inside the condenser. This, as pointed out
earlier, saves tube area, fan coverage and fan power.

8.6 Commercial Designs – Misconception and Shortcomings

We have discussed a simple process where the equipment and process control are
designed in accordance with the underlying physical principles and control objective. It is
interesting to observe some commonly encountered commercial designs, presented
below, of this simple process - designs which display misunderstanding of the physical
principles as well as the control objective.

8.6.1 Condenser Pressure Control

By hot vapour bypass

The ubiquitous hot vapour bypass “technique” makes its way into many refrigeration
units, as shown in Figure 8.9. But it does not work. In Chapter 1.3 we present an in-
depth analysis as to why it does not work. Here, in Figure 8.9, suffice it to say that the
pressure in the condenser or accumulator does not respond at all to the position (opening)
of the control valve in the “hot vapour bypass” line.

177
"Hot vapour bypass"
PD com pressor
(Screw type) Condenser

Oil
Separator
Suction
loader
Oil Cooler
M M
Plenum
PC2 Liquid
Raw gas PC1
KO Drum
LC

Chiller Accum ulator


Gas+liquid

Figure 8.9:
The so-called hot vapour bypass technique for controlling condenser/accumulator
pressure; it actually does not work. The position (opening) of the control valve on the hot
vapour by pass line has no bearing on the pressure in the accumulator or condenser

By manipulating louver opening

One popular control scheme for condenser pressure is by manipulating the louver as
shown in Figure 8.10.

PD com pressor
(Screw type)
Condenser
Oil
Separator

Oil Cooler
M M
Liquid Plenum
PC2
Raw gas KO Drum
PC1
LC

Chiller Accum ulator


Gas+liquid

Figure 8.10:
Wasteful and ineffective way of controlling the pressure in the
condenser/accumulator

Apart from the some undesirable response dynamics (such as the non linear relation
between the louver actuator shaft travel and air flow), this pressure control has one
distinct disadvantage in fan power consumption. This disadvantage is similar to driving
an automobile with both the brake and gas pedals pressed down, simultaneously. Fan
power is wasted through the restriction of the louvers, which restriction is imposed in an
attempt to restrict air flow, with all the fans at full speed!

178
The designer who advocates this control technique is driven by his insistence on
controlling the condenser/accumulator pressure to a fixed target. Hence he is faced with
either this power wasting and ineffective louver manipulation or the expensive hardware
of variable speed motors. This fixed target usually turns out to be the condenser pressure
for the design (i.e. sizing) case; and the design case corresponds to the maximum
refrigeration load at the highest design ambient temperature; this in turns means that in
the design case the condenser pressure is at its highest and the compressor and fans are
run at their maximum powers. So, by virtue of this control scheme, the refrigeration plant
is made to run always in the worst case, namely the sizing case that requires the
maximum compressor and fan powers. As discussed in Section 8.5.3, the right control
scheme is to turn on or off fan motors as needed to ensure a minimum pressure is
exceeded, using louver closure only on extremely cold days when all fans are off.

8.6.2 Compressor Discharge Pressure

In some designs a back pressure control valve is placed on the compressor discharge to
ensure that the compressor always discharges against (or “sees”) a constant (high)
pressure regardless of the prevailing pressure in the condenser, which could be “all over
the map” depending on the ambient temperature. The back pressure controller at the
compressor discharge is in addition to and independent of whatever pressure control
scheme the designer may have for the condenser. Perhaps the designer here is overly
biased by the dynamic characteristics of a centrifugal or axial compressor, which can
only discharge against a very narrow range of pressure at a constant rotational speed; he
may be unaware that a PD compressor is used in this service expressly to take advantage
of the widely varying pressure expected in an air-cooled condenser, against which
pressure the compressor is to discharge. (For discussions on the dynamic characteristics -
also know as performance characteristics - of the PD, centrifugal and axial compressors,
the reader is referred to technical texts that abound on these machines.)

Following are some very interesting arguments made in defence of this “fixed discharge
pressure approach”:

Constant speed motor argument

The compressor has a constant speed motor; therefore the power consumption is constant.
This talk about riding on a lower discharge pressure, even if the compressor is capable of
it, to save power is nonsense, the argument goes. Well, the drive shaft of a motor turning
at constant speed delivers a shaft power that depends on the compressor load. For a given
flow and suction pressure, the higher the discharge pressure the higher is this load on the
compressor. This higher load means higher power delivered by the drive shaft, which,
though turning at constant speed, has to turn against a higher torque, since power =
torque x speed. If this simple relation proves too hard to understand, one can always look
at the ammeter on the motor control panel to see that indeed the motor, always running at
constant speed, does indeed draw a lower current (hence a lower power, at constant
voltage) at a lower compressor discharge pressure.

179
Compressor oil argument

The use of the screw compressor with circulating liquid seal is quite common as the PD
compressor for this service. A schematic of oil circulation is shown in Figure 8.11. (The
“oil” is not a mineral but a synthetic liquid. It is quite often not even an oil but a glycol.)
The primary function of this oil is to act as a dynamic seal between the turning screw and
the fixed components in the compressor casing. This seal allows the set of turning screws
to function like a multistage compressor. The oil also acts as a coolant absorbing the heat
of compression. Lastly, it also lubricates. The compressor discharges into a separator, in
which the oil is separated and cooled and recycled to the suction; the compressed gas
proceeds from the oil separator to the condenser. It is seen that, at a given suction
pressure, a minimum discharge pressure is required to maintain the recirculation of this
oil. Well, there is no need to impose another independent back pressure against the
compressor discharge. All we need to ensure is that the minimum pressure maintained in
the accumulator, on which the compressor discharge should be allowed to ride freely, is
adequate to ensure the delivery of refrigerant to the chiller(s) as well as maintaining this
oil recirculation. See Section 8.5.3 on the control scheme to assure this minimum
pressure.

To Condenser
PD Com pressor
(screw type)

PC3

PD
PS
Oil Separator This discharge rides
Oil on pressure P A in the
circulating refrigerant accumulator

Oil Cooler

From KO Drum/Chiller

Figure 8.11:
Screw compressor oil cooling and circulation:PD > PS sufficiently to maintain
oil circulation; this requirement is met by the prevailing pressure P A in the
accumulator if PD is allowed to ride freely on PA. The pressure control shown
in the cloud is unnecessary and wasteful of compression power.

Design parameter argument

The compressor is sized on the highest demand case: highest refrigeration duty coupled
with highest condensing temperature. At the highest condensing temperature of 43 °C,
say, the bubble point pressure (of propane) is 1471 kPa (a); this is the highest pressure

180
against which the compressor has to discharge. The highest demand case therefore
defines the size and power of the compressor (as well as those of the condenser). The
advocate of "independent” pressure control – i.e. holding a fixed back pressure of 1471
kPa (a) against the discharge, regardless of what the condenser pressure happens to be or
how it is controlled - interprets the discharge pressure corresponding to the highest
demand case as the one at which the compressor has to operate constantly, because that is
the design intent! So, he provides the control to enable this operation, namely a pressure
control valve to hold back 1471 kPa (a) against the compressor discharge, to fulfill what
he believes to be the design intent. Imagine the waste (power and wear and tear) of
discharging against this highest pressure, while a low pressure of 350 kPa (a) may prevail
in the condenser on a cool day, and on this low pressure the compressor discharge should
be allowed to ride freely to save power substantially!

8.6.3 Accumulator Level Control

A few designers insist on having a level control for the liquid held in the accumulator
drum. The conviction of such designers is probably inspired by the importance of level
control in many situations. The reflux drum of a fractionator, for example, has a level
control to hold the level in the drum fixed. The consequence would be dire if that level
were lost: gas would blow into the overhead liquid product line; the reflux pump would
run dry and blow its seals; and a major process upset or fire would ensue. So, argue these
designers, it is just as important here not to lose the level in the accumulator.
PD com pressor
(Screw type)
Condenser
Oil
Separator

Oil Cooler
M M
Liquid Plenum
PC2
Raw gas KO Drum
PC1
LC LC1

Chiller Accum ulator


Gas+liquid

Figure 8.12:
Misconceived accumulator level control (control loop LC1 in the cloud).
If its set point is higher than the actual (fixed) level in the accumulator, LC1 will
drive its control valve to close starving the chiller and compressor of refrigerant.

Well, the refrigerant is in a closed circuit. The circuit holds a fixed inventory of
refrigerant (vapour + liquid). During operation, the level in the accumulator is fixed as
determined by the fixed inventory in the system. There is nothing any control can do
about the level in the accumulator. With a level control in action as shown in Figure 8.12,
only two things can happen: the level control valve is driven to the fully open position

181
and remains at this position, if the set point of the level controller happens to be lower
than the actual fixed level established in the accumulator; or, it is driven toward the fully
closed position, if the set point of the level controller happens to be higher than the actual
fixed level in the accumulator. The former eventuality has no effect on the process and
probably goes unnoticed. The latter will trigger a major process upset. The refrigerant
users and the compressor will all be starved of refrigerant, while this ill-suited level
controller is trying in vain to increase the fixed liquid level in the accumulator to its set
point, by driving its control valve to close completely!

To be sure, the accumulator is fitted with a level gauge to indicate that there is adequate
refrigerant inventory in the system. The refrigerant inventory depletes over time due to
instrument and equipment purging and other maintenance activities; the refrigerant is
topped up occasionally, if necessary, to restore the system to the desired inventory.
(Much in the same way, we don’t control the motor oil level in an automobile – we can’t
because it is fixed - we just watch it and top it if necessary.) If the inventory depletes
faster than expected, it indicates a leak, which must then be investigated.

8.6.4 Pre-saturation of Refrigerant Vapour Entering the Condenser

Very few designs come with the ex situ pre-saturation of the superheated compressor
discharge. In the absence of such pre-saturation, desuperheating will have to be provided
in the condenser. Not only does this lead to greater tube area (hence greater fan coverage
and fan power), it also forces a multi-pass arrangement for the tube bundle in order to
induce an adequate vapour velocity in the tubes to achieve an acceptable heat transfer
coefficient to desuperheat the vapour. The horizontal tubes in a multi-pass bundle offer
very poor drainage, as the liquid in each tube beyond the desuperheating area has to build
up a hydraulic gradient in order to drain. The build-up of this hydraulic gradient in a
horizontal tube causes it to be partially flooded thus losing direct heat transfer area.

Some designers argue against the ex situ pre-saturation of the condenser inlet as follows:

Increased complexity due the need for a circulation pump, circulating line, spray nozzle
and control.

This is a thoughtful argument, but invalid. More elements are introduced into the system.
But for the low flow required for desuperheating the low-head pump usually turns to be a
small in-line pump. In most applications, the circulation line seldom needs to exceed 2
inch diameter in hard piping or ½ inch in tubing. There is no need for automated flow
control; the required flow is set once and left fixed. (See Section 8.5.3 on process
control.) This addition of the hardware and the simplicity of its operation are well
justified by the savings in fan coverage, fan power and a smaller condenser.

182
There is no liquid build-up problem in the horizontal and multi-pass tubes.

This is wishful thinking, inspired by the workings of a partial condenser. For partial
condenser, the tubes are configured to obtain high velocities to sustain mist flow in the
tubes. The mist flow regime is the most efficient for condensation, because the liquid
formed is suspended as mist in the flowing vapour, thus leaving the tube surface
relatively liquid-free and accessible for direct contact with the condensing vapour. The
required vapour velocity to sustain mist flow leads to the multi-pass arrangement. Ideally
the whole flow path, from inlet to the exchanger to the outlet, should be kept in the mist
flow regime. Quite often this is not feasible; the run of tube length toward the outlet may
not have enough vapour flow in it and the mist flow breaks down into stratified flow. In
this run of the tube, however, there should still be enough sweeping action of the remnant
vapour flow to prevent significant liquid hold-up in the tube.

However, in a total condenser such as we have here, we cannot establish a mist flow
regime, not even by design; except, perhaps, fortuitously near the inlet of the exchanger,
where there happens to be high enough vapour velocity to sustain mist flow for a short
run. In a multi-pass tube arrangement, the lower rows of tubes, which have to
accommodate the cumulative liquid loads from the higher rows, are therefore severely
flooded. In these lower rows, which are closer to the outlet, there is hardly any vapour
sweep to alleviate the bad situation created by flooding. As a matter of fact, the vapour
flow at the outlet is exactly zero, because it is a total condenser.

Apart from the benefit of lower tube area, lower fan coverage and lower fan power, ex
situ pre-saturation eliminates the competing mechanism of desuperheating in the
condenser. With the elimination of the need to desuperheat in the condenser, the
condenser can now be configured as a simple single-pass tube bundle sloped and
pressure-equalized for good liquid drainage.

Designers who use a multi-pass bundle, mistakenly, for the total condenser do so for one
or more of the following reasons:

Failure to recognize that the condensing coefficient is practically independent of vapour


velocity.

The failure to recognize the independence of the condensing coefficient (i.e. film
coefficient of vapour condensing) from vapour velocity does not arise from the lack of
data or literature on the subject. Rather the designer is misled by the calculation
procedures commonly adopted for the design and sizing of a condenser. Most if not all
design and sizing procedures, including those automated and packaged into a program or
commercial software, make no distinction between a partial and total condenser. These
procedures may vary in details, but they all have one thing in common: the designer has
to ensure that, in configuring the tube bundle, a certain minimum vapour velocity must be
maintained in the tube; specifically, this minimum refers to the arithmetic average of the
vapour velocity at the condenser inlet and that at the outlet.

183
One can appreciate the need for this minimum velocity for the partial condenser, to
sustain a mist flow regime in the tube. To be sure, a higher vapour velocity directionally
improves the condensing coefficient; it does so by facilitating the dislodging of liquid
formed on the tube surface. This improvement is only marginal and is not the reason
behind the minimum velocity requirement – which reason is to sustain mist flow. Mist
flow need not, and cannot, be maintained in total condensation. Whether or not they are
cognizant of the true reason behind the minimum velocity requirement, many designers
believe that the minimum vapour velocity is to achieve an acceptable condensing
coefficient. This mistaken belief compels them to adopt a multi-pass tube bundle to meet
the minimum velocity requirement, for the partial as well as the total condenser.

Over reliance on computer software

Most commercial programs are equipped with good and extensive data base, good
correlations and calculation methods, supported by literature and research, and a stable
and robust numerical process. Warning messages are given in the output against anything
untoward or out of bound, which give considerable comfort to the designer; but quite
often they also engenders complacency. The user is left with this false sense of
reassurance that, if the program does not warn about something, then it must be all right.
The designer should always adopt the caveat emptor position: “buyer beware”. These
programs are tools he uses at his own risk. It is unlikely that the program will give a
warning like: use of multi-pass bundle results in stratified flow and severe flooding;
consider other options. The user should examine the program closely and ask questions,
such as:

• Does the program distinguish among the various two-phase flow regimes in the
condensing tubes, let alone recognize the effects of these regimes on sizing and
warn the user about it? In particular, does the program distinguish and quantify
flooding? Does stratified flow regime occur in the tube? Given a liquid loading
(liquid flow rate) in a horizontal tube, it is quite straight forward to calculate the
hydraulic gradient (liquid hold-up) that must prevail in the tube in order for it to
drain. For a 4 meter long tube, for example, it does not take much liquid loading
to flood the tube substantially. Is the effective loss of condensing surface due to
flooding adequately accounted for in the program? More often than not it is not
adequately or not at all accounted for by the calculations in the software. When
1/3 of a tube area is flooded, we practically lose 1/3 of the heat transfer area in
that condenser tube. The calculations in the software might proceed as if there
were no loss of surface area due to flooding. Because of the temperature gradient
between the liquid surface and the cold tube wall, heat given off by the vapour
condensing on the liquid surface is removed by conduction through the flooding
liquid layer. This conductive rate is very much lower than the vapour condensing
coefficient; therefore, the effective heat transfer surface is implicitly overstated by
the calculation methods that ignore the insulating effect of the flooding liquid.

184
• How is the condensing coefficient calculated in the software? Is it reasonable?
Check the magnitude of condensing coefficient used in the program for sizing
against the typical ones published in some texts or manuals. Better still, perform a
sanity check by calculating the condensing coefficient independently by hand.

8.7 Ammonia versus Propane as Refrigerant

Propane is used almost exclusively as a refrigerant in gas processing (hydrocarbon dew


point correction) in Alberta, Canada; but occasionally ammonia is encountered. When
ammonia is used in this application, it is almost sure to be a mistake. (The only
conceivable situation in which ammonia is used is one in which the use of propane is
prohibited or prohibitively expensive compared to ammonia – a situation that rarely if
ever occurs.) At the prudent compressor suction pressure of 110 kPa (a) (i.e. 10 kPa
above the atmospheric), ammonia boils at -31 °C, whereas propane boils at -40 °C.
However, we note that for propane to boil at -31 °C the corresponding pressure is 159
kPa (a). To make a comparison of ammonia and propane on an equal footing, we apply
the same design parameters to the mechanical refrigeration: refrigerant boiling at of -31
°C in the chiller and condensing at 43 °C in the condenser as shown in Table 8.1.

Table 8.1:
Comparison of Propane and Ammonia
Refrigeration duty: 100 ton (or 293.1 kW chiller duty)
Condenser temperature: 43 °C

Chiller @ -31 °C Chiller @ -40°C(1)


(equal footing) (propane only)
Propane Ammonia Propane
Refrigerant flow, kgmole/h 57.8 44.7 56.6
Latent heat in chiller, kJ/kgmole 18245.6 23624.8 18629.4
Compressor suction, kPa (a) 159 110 110
Compressor discharge, kPa (a) 1471 1678 1471
Compressor shaft power, kW (2) 98.7 123.6 112.9
(3)
Condenser duty, kW 391.7 416.6 406.0

(1) Enhanced liquid recovery. Chiller temperature of -40 °C is not possible with
ammonia without reducing the compressor suction pressure to 65 kPa (a), which
is sub-atmospheric and therefore unacceptable.
(2) At adiabatic efficiency of 75 %.
(3) No inter-stage cooling of compressed vapour.

It is clear that on the equal footing the compression power and condenser size are
significantly smaller for propane compared to ammonia despite the fact that the mass
flow rate of ammonia is lower for the refrigeration duty. In addition to the foregoing

185
advantage of propane over ammonia, we have the following location specific
considerations:

• By comparing the boiling points of -40 °C and -31°C, respectively, of propane


and ammonia at 110 kPa (a), it is clear that with the use of propane refrigerant the
recovery of light hydrocarbon liquid from the raw gas is enhanced. While the
chiller temperature of -31 °C may be low enough to produce gas that meets the
pipeline hydrocarbon dew point target, the enhanced liquid recovery achieved
with use of a chiller temperature of -40 °C usually represents a revenue increase
for the operator. In the regime of product values in many locations, light
hydrocarbons (ethane, propane, butanes) are worth substantially more as
segregated liquid product streams than as components left in gas product stream.

• In all gas processing facilities, light liquid hydrocarbons (propane, butanes) are
already present in the processing scheme. The use of propane as refrigerant does
not constitute an additional material handling requirement. In contrast, the use of
ammonia introduces an additional material handling requirement (assuming that
ammonia is not used elsewhere in the facilities.)

• In many gas processing locations refrigerant grade propane is readily available in


relatively close proximity. In contrast, the use of ammonia in these locations
usually entails transportation of this hazardous material over long distance.

In Table 8.1, it is also shown that the propane scheme (chiller temperature of -40 °C and
compressor suction pressure of 110 kPa (a)) still maintains an advantage, in terms of
compression power and condenser duty, over the ammonia scheme (chiller temperature
of -31 °C and compressor suction at 110 kPa(a)). Here is a rare win-win moment: the
propane scheme gives better liquid recovery for lower power consumption and a smaller
condenser. This is not too good to be true; it simply is true. The truth is largely due to the
fact that, to condense at 43 °C, ammonia requires a pressure of 1678 kPa (a) whereas
propane requires only 1471 kPa (a); the lower condenser pressure for propane results in
lower compression power, in spite of the higher circulation rate of propane for the same
refrigeration duty.

One can appreciate that the designer may not be aware of the location specific
considerations that favour the use of propane as refrigerant, such as accessibility, material
handling issues, and the product value regime of the liquid products obtained. But he
should be aware of the advantage of propane over ammonia on purely technical bases
such as shown in Table 8.1. So, in gas processing for hydrocarbon dew point correction,
it is a mistake to use ammonia as refrigerant. Ammonia is used only because the designer
happens to be “familiar” with it – familiar in the sense that he has designed refrigeration
systems with ammonia as refrigerant; and it does not occur to him to consider propane as
a refrigerant. One can understand the comfort of working with the familiar. Some
designers, however, take this “familiarity” too far and pass it for “expertise”. In one
instance, a designer was asked to explain the choice of ammonia over propane; he
enunciated to a group of design reviewers that, according to his experience, ammonia is a

186
more efficient refrigerant than propane – he actually said so with a straight face! By the
sheer force of the expertise and experience, real or perceived, attributed to him, a
designer can get away with such misinformation, usually but not always, because once in
a while he may inadvertently come across a reviewer who is somewhat more analytical
and less credulous.

187
Appendix I:

Calculation Routine for Steam Desuperheater Control

Set Point for Desuperheating Water (BFW) Flow Control

FY
34

FI TI PI 41-H-1
34 34 34
Superheated Steam Desuperheater
Saturated Steam
Pressure P kPa (g) SP Pressure P kPa (g)
o
Temperature T C
FC Enthalpy H SAT kJ/kg
Enthalpy H SH kJ/kg 29
Flow F SH kg/h

BFW
o
Temperature T BFW C
Enthalpy H BFW kJ/kg

Figure 1:
Control scheme for steam desuperheating to saturation

Routine for Calculator FY-34

1. From inputs of PI-34, TI-34 and FI-34, calculate F SH (kg/h), the pressure- and
temperature-compensated mass flow rate of superheated steam.

The computer for process control (DCS) comes with a data base for steam. It also
comes with programmable software which can be programmed to access these
data and perform calculations with them. Using this programmable software:

2. Determine enthalpy H SH (kJ/kg) of superheated at measured pressure and


temperature as given by PI-34 and TI-34.

3. Determine enthalpy H SAT (kJ/kg) of saturated steam at measured pressure as given


by PI-34.

4. Determine the enthalpy of BFW (boiler feedwater) at temperature T BFW . (This


temperature may be measured continuously and the signal is input to the
calculation; or, in the absence of such continuous measurement, a constant
nominal T BFW may be used without incurring significant error in the final result,
namely F BFW )

188
5. By applying the first law (energy conservation) to the desuperheater, determine
the BFW flow F BFW (kg/h) required to saturate the incoming superheated steam
as:

H SH − H SAT
FBFW = × FSH
H SAT − H BFW

6. Multiply F BFW by 1.03 to obtain the output of FY-34, i.e. FY-34 OUTPUT = 1.03
x F BFW (kg/h). This output becomes the set point for FC-29 when cascaded. See
Figure 1.

Notes:

• The 1.03 bias should take care of errors inherent in measurement and calculation
(especially of enthalpy values). Also, it is better to slightly overspray with BFW
to guarantee saturation than to spray not enough and miss saturation. A small
amount of moisture in the saturated product steam has no detrimental effect on the
process, whereas superheated (i.e. insufficiently desuperheated) steam
commandeers inordinate amount of exchanger tube area and thus causes shortfall
in exchanger duty. Based on the knowledge of the actual accuracy of
measurement of the instrument, acquired by operating experience, this bias can be
increased if necessary to ensure saturation.

189
Appendix II

Principle of Launching with Angular Momentum

Figure 1 (a) shows an aggregate of liquid elements each having a certain angular
momentum, all angular momenta being of course in the same direction – i.e. they all spin
one way so that their spins augment and not cancel each other. Which way should we
launch them spinning? The short answer is: it doesn’t matter. One might argue that, in
the case of a horizontal flow area (cross section) situated in the northern hemisphere, they
should be launched to spin clockwise, namely in the same direction as the background
Coriolis force due to the earth’s rotation. But this background Coriolis Effect is
insignificant compared to the angular momentum imparted by the launcher. Since we can
make it spin in one way just as easily as in the other, we might as well launch the liquid
spinning clockwise (in the case of a horizontal flow area situated in the northern
hemisphere), for the sake of finesse and as a token of appreciation for Mother Nature’s
gesture of a helping hand.

Of course, what we see is not a myriad of spinning liquid elements covering the circular
cross section, as Figure 1(a) suggests. Instead, what manifests physically is the aggregate
of the individual angular momenta, as depicted in Figure 1 (b), which shows a continuous
body of liquid rotating about the centre of the circle. The centrifugal force of rotation
spreads the incoming liquid evenly to fill the entire circular cross section; this tendency to
spread outward is supplemented by the superimposed radial flow of the liquid as it leaves
the launcher. The superimposed radial flow is due to the successive elements being
launched; the elements being launched push those launched ahead of them radially
outward toward the circumference of the vessel.

(1)
Launcher
(b) Net result of this launching is a
(a) Myriad of tiny spinning
continuous rotating liquid body
liquid elements launched
continuously (1) Launcher contrived to impart
angular momentum to the liquid launched

Figure 1:
Manifestation of angular momentum

190
In a real liquid the rotating body of liquid is subject to retarding viscous force. This
viscous force retards rotation as the liquid moves toward the circumference of the vessel.
The “trick” therefore is to impart to the launched liquid at high enough rate of angular
momentum so that every element of the launched liquid makes its way from the launcher
at the centre to the circumference without being ground to a halt by viscous retardation.

By dimensional analysis, a dimensionless number N b is obtained:

ρΓ
Nb = , where:
µ 2D
Nb : broadcast number (for want of a better name and alluding to the grass seed
broadcaster one uses to seed the lawn)
ρ : density of the liquid-liquid mixture
µ : viscosity of the liquid-liquid mixture
D : diameter of the circular cross section of the vessel
Γ : rate at which angular momentum is imparted to the liquid being launched, i.e.
torque exerted by the launcher on the liquid being launched.

For the launched liquid to spread out to cover the entire circular cross section of the
vessel, a broadcast number N b ≥ 200,000 is required. For most liquids of low viscosity
(water, many oil-water emulsions) a broadcast number of this magnitude is quite easy to
obtain with a simple design using only the force of gravity to impart the required spin
(angular momentum) to the liquid. Appendix II shows the design considerations and
sizing for two gravity-driven launchers, as well as the derivation of the broadcast number
by dimensional analysis. Otherwise, an external motive power may be required to apply
sufficient torque to the launched liquid in order to surpass this threshold of 200,000 of the
broadcast number N b .

Note that the broadcast number N b is proportional to the density of the liquid launched. If
the liquid contained gases the effective density of the launched fluid would be much
lower, thus diminishing the broadcast number. For the same torque Γ applied, we would
get less broadcasting reach with a gassy liquid. Intuitively, we know this to be true, as it
is much easier to set spinning a volume element of liquid than the same element of liquid
mixed with gas. This detrimental effect of gas to liquid launching is quite apart from the
loss of quiescence (the egg-beater effect) it causes in the liquid body.

191
Appendix III:

Design and Sizing Considerations for Liquid Launcher for Use in a Separator

In this appendix we present a launcher with toroidal nozzles. In this launcher, the liquid is
launched with a tangential velocity with respect to the radius of the launcher. Angular
momentum is imparted to the launched liquid by the drive head set up in down pipe. The
angular momentum is what propels the launched liquid to spread out toward the vessel
wall covering the flow area (vessel cross section) evenly. It should be noted that the
launcher diameter is small compared to the vessel diameter.

1 Launcher with Toroidal Nozzles

A
Launcher radius r Toroidal tube centre line (typ)
(semi-circle of diameter = r )

Inlet distributor
(for nozzles)

Dow n pipe
V Inlet distributor
Dow n pipe
(for nozzles)
V
Launched velocity
A (tangential) Nozzle (typ) -
Nozzle (typ) - truncated toroidal tube
truncated toroidal tube
View A - A
Figure 1
Launcher with toroidal nozzles

ρΓ
: broadcast number dimensionless
µ 2D
ρ : liquid (emulsion) density kg/m3

Γ = m Vr : torque applied to launched liquid Nm

µ : liquid (emulsion) viscosity Ns/m2 (or Pa.s)

D : tank (vessel) diameter m

m : liquid mass flow rate (total liquid) kg/s

192
V : launched velocity m/s

r : launcher radius m

For adequate broadcast of the launched liquid to cover the whole circular cross section of
the tank (vessel), the broadcast number ρΓ µ 2 D must be greater than 200,000. (See
Section 4 of this appendix for the derivation of the broadcast number). For low viscosity
fluids, water, oil-water emulsions, this magnitude of the broadcast number is easily
attainable with gravity drive alone. An example of the design and sizing calculations for
the launcher with toroidal nozzles is given in Section 5 of this appendix.

Feed

Gas breakout
pan

Descent
Drive head
velocity u
Dow n pipe available
L hD Dow npipe

Oil surface

Oil-w ater
interface o 1  u2 
Head loss 90 =  
direction change 2  2 g 
Toroidal nozzle

Figure 2:
Launcher with toroidal nozzles: deployment in a vertical cylindrical vessel (tank)

Energy balance –drive head required

d : downcomer pipe ID m

u : velocity in down pipe m/s

1V2
(1) : kinetic head of launched liquid m
2 g

1 u2
(2) : head loss of 90 ° turn in down pipe m
2 g

193
L u2
hF 1 = 4 f (3) : friction loss in downcomer pipe length L m
d 2g

f : Fanning friction factor in down pipe dimensionless

1V2
hF 2 = (4) : entry loss to toroidal nozzle m
4 g
hD* : gravity drive head minimum required: m

hD* = (1) + (2) + (3) + (4)

For applications with water and oil-water emulsion, a reasonable gravity drive head
available hD ≥ hD* can be obtained with a velocity u in the downcomer pipe of about 1
m/s or less.

2 Augmented Drive Head

PC
SP
Gas

Feed LC

Annular Degas Drum


oil trough
Liquid-liquid
Interface

Launcher
Separator
(vertical cylindrical tank)

Figure 3:
Drive head for launcher augmented by Degas Drum pressure

In the arrangement shown in Figure 2, the gas breakout pan is situated inside the vessel.
The drive head available hD is then the liquid head that builds up in the downcomer pipe
above the liquid surface in the vessel. In this arrangement, the launching is purely
gravity-driven. This is adequate for most applications with oil-water emulsions. (Note
that the broadcast number is inversely proportional to viscosity squared; the low viscosity
of water contributes to a high broadcast number.) However the drive head can be
augmented for liquid mixtures with much higher viscosity than oil-water emulsions, as
shown in the arrangement in Figure 3.

194
Note that the pressure in the degas drum is higher than that in the separator (by
approximately the level difference between the degas drum and the separator). Some
residual degassing will occur at the launcher in the separator, but this should have minor
impact on launching.

3 Deployment of Launcher in Horizontal Cylindrical Vessel

Liquid surface
Flow

o
90 EL

Launcher (1) Launcher


Launcher axis

View A-A: Flow area


A (truncated circle formed by wetted perimeter and liquid surface)

(1) Launcher axis to be parallel to vessel axis and to pass through controid of flow area

Figure 4:
Deployment of launcher for a horizontal separator; launcher with toroidal nozzle or with twisted tape

In this deployment, the minimum drive head required hD* should also include a fifth term
(5) to account for the head loss in the 90 ° elbow that orients the launcher to broadcast in
a vertical plane – the plane of the flow area for a horizontal separator.

Energy balance –drive head required

d : downcomer pipe ID m

u : velocity in down pipe m/s

1V2
(1) : kinetic head of launched liquid m
2 g

1 u2
(2) : head loss of 90 ° turn in down pipe m
2 g

195
L u2
hF 1 = 4 f (3) : friction loss in downcomer pipe length L m
d 2g

f : Fanning friction factor in down pipe dimensionless

1V2
hF 2 = (4) : entry loss to toroidal nozzle m
4 g
hD* : gravity drive head minimum required: m

1  u2 
  (5) :Head loss in the 90 ° elbow m
3  2 g 

hD* = (1) + (2) + (3) + (4) + (5) m

4 Dimensional Analysis to Obtain the Broadcast Number

(2)
Water IN
S

Water outlet
A Launcher
(1) A nozzle
D
h Launcher

(1) Launcher w ith toroidal nozzle; launcher elevation adjustable to vary


residence volime and aspect ratio h /D
(2) Water inlet rate is adjustable

Figure 5:
Set-up of a test tank and key dimensions

For a central launcher in a cylindrical vessel, a study was conducted to find out the rate at
which angular momentum must be imparted to the launched liquid such that the launched
liquid spread out to cover the whole circular section of the vessel. The study used the
technique of dimensional analysis. An illustration of the dimensional analysis is given by
Giles et al (12). The test liquid was water; the vessel used was a transparent cylindrical

196
tank of 1000 mm (diameter) by 2000 mm height. The launcher with toroidal nozzles was
used for the study. (See Figure 7.)

We postulate the following plausible function for dimensional analysis:

φ (s, Γ, ρ , µ , D, h ) = 0

The symbols in this function are as defined in Section 1of this appendix, with two new
ones:

s: spread of the launched liquid, i.e. the diameter of dotted-lined circle in


Figure 7, which defines the limit to which the rotating fluid reaches before all angular
momentum is lost to viscous retardation.

h: elevation of the launcher from the tank bottom

The rate of angular momentum is equal to the torque Γ exerted by the launcher on the
liquid leaving it. For the launcher with toroidal nozzles it is given by Γ = m Vr . (See
Section1 above.)

The dimensions of the quantities appearing in the postulated function are:

[s ] = L
[Γ] = force x length = ML2T-2
[ρ ] = ML-3
[µ ] = ML-1T-1
[D] = L
[h] = L
We have 3 counts of dimensions M, L and T (mass, length and time) and 6 quantities. By
the theorem of dimensional analysis we must have 3 (= 6 – 3) dimensionless numbers.
We use the Buckingham π method to obtain these three dimensionless numbers; see
Giles, R. V. et al (12). For the first dimensionless number π 1 , we choose 3
quantities Γ, ρ , µ , each raised to an unknown power, and another quantity D , called
adjunct quantity, raised to an arbitrary** constant power -1, in accordance with the
Buckingham π method, as:

π 1 = Γ m ρ n µ l D −1 ; or:
(
0 = ML2 T − 2 ) (ML ) (ML
m −3 n −1
)l
T −1 L-1 = M m + n +l L2 m −3n −l −1T −2 m −l

Equating powers of M, L and T to zero, we have:

m+n+l = 0
2m − 3n − l − 1 = 0

197
− 2m − l = 0

From this we have: m = 1, n = 1, l = −2 ; so the first dimensionless number is:


ρΓ
π 1 = 2 , which we call broadcast number
µ D

The remaining two dimensionless numbers π 2 , π 3 can be obtained from the remaining
quantities s, h, D by inspection. (Note that D has been picked only as an adjunct quantity
to obtain the first dimensionless number π 1 . This means, in the protocol of the
Buckingham π method, it is still available as one of the 3 remaining quantities from
which we can obtain two more dimensionless numbers.) We judiciously choose them to
be:

s
π2 = , which we call spread number; and
D
h
π 3 = , which we call the aspect ratio of the residence volume
D

In accordance with the principle of dimensional analysis, we can establish the following
functional relations of the variable governing the operation of the launcher:

a b
s  ρΓ   h 
= k  2    ; where (1)
D µ D D

k, a and b are constants which can be determined empirically by measurement made with
a physical set-up that models the launching operation, such as the tank test shown in
Figure 5.

For our purpose, we need not obtain these constants; instead, we need only determine, by
measurement, the threshold broadcast number ρΓ µ 2 D at which the spread number s D
attains the value of 1, at a given aspect ratio h D . This is tantamount to finding out, given
the liquid properties and the geometry of the physical set-up, the rate at which angular
must be imparted to the launched liquid so that it spreads to cover completely the cross-
section of the vessel. From the measurements obtained from the tank test, the threshold
ρΓ µ 2 D was determined to be 200,000. At the tank test, the spread of the launched water
was made visible with a pulse of tracer injected to the water inlet into the tank. It goes
without saying that at an operating broadcast number exceeding the threshold of 200,000,
we are assured of complete spread of the launched liquid. The measurements also
indicated that the spread ratio s D is independent of the aspect ratio h D (within the
practical range of the aspect ratio h D of a cylindrical separator); in other words, b ≈ 0 in
the relation given by (1).

198
For liquids of relatively low viscosity (< 10 mPas), this threshold of 200,000 can be
easily achieved by a drive head provided by gravity alone, i.e. without any external
power assist.

_______________________________________________________________________

**It should be noted that any real number can be chosen as the power to which the
adjunct quantity D is to be raised. Usually we choose 1 or -1. Had we chosen 1 (instead
of -1) we would have found the first dimensionless number to be µ 2 D ρΓ , i.e. the
reciprocal of what we found with power -1 for D. It is clear that this has no bearing on the
functional relation we seek of the 6 quantities, as embodied in (1). In general, for a given
phenomenon under study, the arbitrary constant power we choose for an adjunct quantity
has no bearing on the functional relation of the quantities we seek to establish.
_______________________________________________________________________

199
5. Examples of Launcher Design Calculations

Emulsion viscosity = 20 mPas

ρΓ
ρΓ
Broadcast number Dimensionless 508,000
µ 2D
3
ρ Liquid density kg/m 1000
1.016
Γ = mvr , torque imparted to incoming fluid Nm

2
µ Liquid viscosity N s/m (Pa.s) 0.020

5.00
D Tank diameter m

m Liquid mass flow rate (total launcher) kg/s 5.00

v Launched velocity m/s 1.00

r Launcher radius m 0.203

For adequate broadcast of the launched liquid to cover the whole cross section of the tank,
the broadcast number must be greater than 200,000. [For low viscosity fluids, a broadcast
number of this magnitude is achieved easily.]

L downcomer pipe length m 5.00


d diameter of downcomer pipe m 0.102
u velocity in downcomer m/s 0.617
2
1V
kinetic head of launched liquid m 0.025 (1)
2 g
1 u2
head loss in 90 degree turn in down pipe m 0.010 (2)
4 g
ε Absolute roughness of pipe 0.0018 inch
d Pipe ID 4.000 inch
NR Reynolds number 3,133

fc Fanning frcition factor 0.0108

L u2 friction loos in downcomer pipe m 0.041 (3)


4f
d 2g

1V2
= entry loss to toroidal nozzle m 0.025 (4)
4 g

hD* minimum drive head req'd m 0.102

hD* = (1)+(2)+(3)+(4)

200
Appendix IV

Bernoulli Equation from the First Law

The first law of thermodynamics (law of energy conservation) gives the following
equation. We show the construction of this equation in Chapter 7: Thermodynamics – An
Overview.

 u2   u2  d
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g  = q − w S − (mC U C ) (1)
 2   2  dt

m 1 , U 1 , P1 , V1 , u1 , z1
Material IN
1 m 2 , U 2 , P2 , V2 , u 2 , z 2
Material OUT

Surroundings System

\
Heat q = 0 Work w
s = 0
mC : material holdup inside system
U C : internal energy of material inside system

Figure 1:
A steady-state system; only exchange of material with the
surroundings; no exchanges of heat and work.

The system shown in Figure 1 is steady-state; and there is, by stipulation, no exchange of
heat and work between it and the surroundings: q = 0 and w S = 0. Material enters the
system at the rate of m 1 and leaves it at m 2 ; the states and properties of the material at the
point of entry (point 1) and the point of exit (point 2) are as marked in the figure.
Applying the first law equation (1) to this system, we have:

q = 0, w S = 0 (by stipulation, no exchanges of heat and work);

d (mCU C ) dt = 0, and m 1 = m 2 (steady-state)

 u 22   u2 
U 2 + P2V2 + + z 2 g  − (U 1 + P1V 1 ) + 1 + z1 g  = 0 (2)
 2   2 

Note that (2) is obtained from (1) by specializing it to a particular system; nonetheless (2)
is still exact, because in its derivation no terms have been considered insignificant and

201
deleted. The specific volume V is just the reciprocal of density ρ, which we assume to be
constant for our system. We re-write (2) by multiplying it through by ρ and re-arranging:

u12 u2
P1 + ρ + ρz1 g = P2 + ρ 2 + ρz 2 g + ρ (U 2 − U 1 ) (3)
2 2

Each term in (3) has the unit of energy per volume. All these terms, except ρ (U 2 − U 1 ) ,
are called heads: pressure heads (for P 1 and P 2 ); static heads (for ρz 1 g and ρz 2 g); and
velocity (or kinetic) heads for ( ρ u12 2 and ρ u 22 2 ). It is customary to divide (3) through
by ρg so that each term now has the dimension of length (or height):

P1 u12 P2 u 22 (U − U 1 ) (4)
+ + z1 = + + z2 + 2
ρg 2 g ρg 2 g g

We recognize (4) as what is known as the Bernoulli equation encountered in fluid


mechanics, which is usually written as:

P1 u12 P u2
+ + z1 = 2 + 2 + z 2 + hF (Bernoulli equation) (5)
ρg 2 g ρg 2 g

Observations:

Gain in internal energy in a dissipative process

The term h F in the Bernoulli equation is the friction loss; it is dissipative, non-recoverable
loss. This loss is dissipated as heat. By purely mechanical consideration, we know h F
must be a positive quantity, since friction is positive loss to the system. We know that
under the ideal condition of zero friction loss the sum of the heads on the left of (5) will
manifest as, or transform into, the sum of all the heads on the right, with h F = 0.
However, under real conditions in which loss is inevitable, not all of the total heads on
the left transforms into the heads on the right, and the portion of it that is not so
transformed is lost, as represented by h F , which, as it is written in (5), must therefore be a
positive quantity.

Since hF = (U 2 − U 1 ) g > 0 , we must therefore have U 2 > U 1 . The material actually gains
internal energy U by flowing through the real dissipative system. It has more internal
energy (at point 2) than that it started with (at point 1) due to the temperature rise brought
about by the friction loss dissipated as heat.

202
Loss and transformation

We need to be careful when we talk about energy, head and head loss. Head loss refers to
h F ; it is the part of the sum of inlet heads that is not transformed into the sum of outlet
heads, but lost irreversibly, for ever. A change in head is not a head loss. For example the
( )
change u 22 − u12 2 g is neither a gain nor loss; it is a transformation. Similarly, the
changes (P2 − P1 ) ρg and z 2 − z1 are transformations, not head losses (or head gains).
The sum of these three changes is zero for an ideal (no friction) system; for anything real
this sum is non-zero and equal to -h F .

Application of the Bernoulli Equation to the Skimmer

LTT
Inlet LG
Gas /
LW h A : overflow weir height
vapour
y , oil layer thickness hL : oil weir height
LS
1
2
H : crest heght
H D B : overflow weir width
hW hL
hA Vo : approach velocity in crest
Underflow
(overflow horizontal velocity)
(a) Major vessel dimensions
y : oil layer thickness
Oil Aqueous
(Amine) Vo2
: overflow kinetic head
Oil w eir (trough edge)A 2g

Oil trough
Vo
H
H

hL Underflow
w eir
Upflow

B
hU
hA
Underflow hA
WUP Overflow w eir
View A - A
A
(b) Underflow weir and upflow channel width (c) Overflow weir height and width

Figure 2:
Skimmer: Dimensions and nomenclature

Application of the Bernoulli equation to the skimmer is straight forward. Figure 2 shows
points 1 and 2 that will be used in our analysis. In this application, we make the following
assumptions and substitutions to equation (3); the substitutions are made to accommodate
the particular nomenclature and physical set-up we have for the skimmer:

203
P 1 = P 2 (the vessel is under one single inside pressure)
u 1 = 0 (insignificant velocity head in the liquid body in the residence compartment)
u 2 = V o (amine velocity over the crest)
ρz 2 g = ρ A (h A + H )g ρz1 g = ρ o y + ρ A hW = ρ A (hL − y )g + ρ o yg
u 22 Vo2
ρ = ρA
2 2

ρ (U 2 − U 1 ) = ρ A hF g

Using the above substitutions, equation (3) becomes the design equation for the skimmer:

Vo2
ρ A (hL − y )g + ρ o yg = ρ A (h A + H )g + ρ A + ρ A hF g (6)
2

Notes on the Bernoulli equation

The full expression of the energy conservation equation (Bernoulli equation) is:

V R2 u2 V2
ρ A (hL − y )g + ρ o yg + ρ A − ρ A hFR g − ρ o o = ρ A (h A + H )g + ρ A o + ρ A hF g
2 2 2

(7)

The term (ρ o y + ρ A hw ) , which is equal to ρ A (hL − y ) + ρ o y , is used in the discussion in


Chapter 4: Oil Skimming in a Horizontal Separator and elsewhere to make it more
transparent as to which way the oil-water interface moves to compensate for a change in
the term ρ A (h A + H ) , in particular a change in the amine overflow crest height H during
operation. The third and fourth terms on the left ρ AVR2 2 and ρ A hFR g are, respectively,
the kinetic head and the loss due to the movement of the liquid body in the residence
compartment; the last term on the left ρ o u o2 2 is the kinetic head of the oil overflowing
into the oil trough. These three terms are insignificant even in the context of the delicate
balance in the application to the skimmer; they are therefore knowingly ignored in
constructing the design equation (6) for the skimmer.

204
Appendix V:

Design of Three Phase Separator with Oil Skimming


=========================================
V.1 Design Principle
V.2 Process Dynamics
V.3 Calculation Procedure
V.4 Calculation Flowchart
==========================================

V.1 Design Principle

The vessel dimension (length x diameter) should be such that:

1. The gas space provides enough residence time for the entrained liquid droplets to
fall onto the liquid surface before the gas flows out of the residence compartment;

2. There is enough liquid volume held up in the vessel to meet the required aqueous
phase residence time for the oil droplets to rise to the oil layer, before the aqueous
phase leaves the residence compartment by the underflow weir;

3. An acceptable oil layer thickness y (100 to 120 mm) is maintained;

4. An acceptable crest height H (25 to 75 mm) is established over the aqueous phase
overflow weir; and

5. There is enough surge volume in the aqueous phase surge compartment.

V.2 Process Dynamics

The design equation for the skimmer is the energy conservation (Bernoulli) equation:

 2
 (1)
(ρ o y + ρ A hW ) = ρ A (hA + H ) + ρ A  Vo+ hF 
 2g 

It should be noted that:

1. The higher the overflow weir h A the thinner the oil layer y becomes. This is
obvious because as h A is raised, the oil-aqueous interface has to move up to gain
enough head, given by (ρ o y + ρ A hW ) , in order to compensate for the now higher
head ρ A (h A + H ) sustained by the aqueous phase overflow weir. It moves up
because ρ A is greater than ρ o . In fact h A could be so high that an oil layer cannot
be sustained (i.e. y is reduced to zero). In this situation, the separated oil is swept

205
continuously as a sheen by the aqueous phase overflowing the oil weir into the oil
collection trough. In short, the oil skimming process fails to work – the oil trough
collects a continuous and significant stream of oily aqueous phase instead of
relatively dry skimmed oil! In many applications, this aqueous phase is an
expensive amine solution.

2. For a given h A , the higher the oil density ρ o (due to heavier oil in the feed or due
to the weathering of the resident oil layer) the greater the oil layer thickness
y becomes, and vice versa.

3. The higher the aqueous phase throughput, the smaller the oil layer thickness
y becomes, and vice versa. This is so because at higher throughput the crest
height H increases. The oil-aqueous interface will have to move up so that
enough is gained by (ρ o y + ρ A hW ) to compensate for the higher H in ρ A (h A + H ) .

4. In an oil skimmer application, the oil presence in the feed is incidental and small -
it is in principle zero. The oil layer is maintained for operability and not for
residence time. The gratuitous residence time the oil phase has in the oil layer is
usually in the order of hours or days. The resident oil suffers prolonged
weathering. Therefore weathered oil density should be used instead of the fresh
oil density from the probable sources.

Sensitivity

From the discussion above, it is clear that an acceptable operating oil layer must be
sustained at all times. As suggested by the energy conservation equation, the oil layer
thickness y is sensitive to the aqueous phase overflow weir height h A , the crest
height H and the operating (weathered) oil density ρ o . The operating margin allowed for
the oil-aqueous interface movement is ± 50 mm. This margin is imposed so that, under
extreme operating conditions, the oil layer does not disappears completely (at high
throughputs); or, at the other extreme, grows so thick (at low throughputs) that it deprives
the aqueous phase of its residence time substantially, or worse still, gets dragged through
the underflow into the aqueous phase compartment. In the light of the fore-going, the
design effort for an oil skimmer calls for the following considerations:

1. Because of its gratuitous, indeterminate but very long residence time, the oil layer
held in the residence compartment suffers considerable weathering, which results
in higher oil density ρ o . This higher “weathered” density should be used in the
design calculations.

2. Judicious choice of design oil layer thickness y and corresponding overflow weir
height h A to sustain this thickness;

206
3. Rigorously calculated crest height H ;

4. Because of the narrow operating margin allowed for the movement of the oil-
aqueous interface, the kinetic head Vo2 2 g of the overflowing aqueous phase must
be taken into account. The kinetic head, for example, is in the order of 20 to 25
mm of liquid column – hardly negligible in this balance.

5. The head loss hF (due to friction and flow direction change) incurred in the
aqueous underflow and upflow must be included in the design equation. It must
be kept low and insignificant (< 3 mm) by using appropriate underflow weir
clearance and upflow channel width. It is retained in the calculations as an
indication that we do keep it insignificant, in consideration of the sensitivity of the
oil layer thickness y.

6. The design must be rated to ensure that an acceptable minimum oil layer of
thickness of 25 mm is still sustained at 120 % design throughput.

To underscore the importance of the design effort outlined above, let’s repeat the
sensitivity analysis presented in Chapter 4:

ρA   Vo2 
y= (
 Lh − h A ) − H −  + hF  (2)
ρ A − ρ o   2g 

Taking ρ A , ρ o and (hL − h A ) as parameters, we can take the differential of the preceding
re-arranged Bernoulli equation:

∂y ∂y ∂y
) (
d Vo2 2 g + )
dhF , which expresses
dy = dH + (2a)
∂H (
2
∂ Vo 2 g ∂hF

the variation of y due to changes in H, Vo2 2 g and hF .

Taking the finite difference approximation of (2a), we have:

ρA  Vo2 
∆y =  − ∆ H − ∆ − ∆hF  ; noting that: (2b)
ρ A − ρo  2g 
∂y ∂y ∂y − ρA
= = =
∂H ∂ (Vo 2 g ) ∂hF ρ A − ρ o
2

In (2) and its transform (2b) we have used the substitution ρ o y + ρ A hW


= ρ o y + ρ A (hL − y ) . For want of a better word, let’s call the density ratio ρ A (ρ A − ρ o )
amplification factor. It amplifies the effect on the oil layer thickness y of a change in any

207
of the terms in the square bracket on the right. At ρ A = 1.00, ρ o = 0.80, the amplification
factor is 5. This means a change, say, in H by +4 mm (due to change in throughput)
causes the oil layer thickness to change by (-5 x 4 mm =) -20 mm; in other words, an
increase in the crest height of 4 mm will cause the oil layer to become thinner by 20 mm.
At ρ A = 1.00, ρ o = 0.90, the amplification factor is 10! The call for rigour in the design
effort, therefore, cannot be over-emphasized.

V.3 Calculation Procedure

LTT h A : overflow weir height


Inlet LG hL : oil weir height
Gas /
LW vapour H : crest heght
y , oil layer thickness Vt LS B : overflow weir width
Vo : approach velocity in crest
H D (overflow horizontal velocity)
hW hL
hA y : oil layer thickness
Underflow
Vo2
: overflow kinetic head
2g
(a) Major vessel dimensions Oil Aqueous
(Amine)
Oil w eir (trough edge)A

Oil trough
Vo
H
H

hL Underflow
w eir
Upflow

B
hU
hA
Underflow hA
WUP Overflow w eir
View A - A
A
(b) Underflow weir and upflow channel width (c) Overflow weir height and width

Figure 1:
Skimmer: Dimensions and nomenclature

The design calculations are quite simple in principle. The calculations have an
appearance of complexity due to the fact that iterations, both local and global, are
required to settle on an acceptable set of answers (oil layer thickness, crest height, vessel
slenderness ratio). Adding to this apparent complexity is the use of lengthy equations
(exact expressions) of circular and elliptical geometry for the calculations of flow area
and volume. In the past era of the slide rule and pocket calculator, interpolation of
tabulated data or readings from charts were used for calculations involving circular and

208
elliptical geometry because of “unwieldiness” of the exact expressions. Fortunately
nowadays, thanks to the spreadsheet and other easily accessible user-friendly
programmable software, some of us have learned to abandon this antiquated labour-
intensive practice and use the exact expressions instead.

Given the apparent complexity, the calculation procedure is therefore best expressed with
a flowchart. (See V.4: Calculation Flowchart.) The procedure can be easily implemented
in a spreadsheet. Numbers (1) , (2), etc, as used in the description below of this
calculation procedure are referred to in the calculation flowchart.

(1) Design parameters

These are parameters to be prescribed by the designer:

Liquid volumetric throughput QA


Gas volumetric rate QG
Cut-off liquid droplet diameter in gas phase dm
Aqueous phase density ρA
Gas phase density ρG
Gas phase viscosity µG
Skimmed oil density (weathered) ρo
Required aqueous phase residence time θA
Required aqueous phase surge time θS
(in the surge compartment)

(2) Initial selection of vessel diameter and liquid height

Refer to Figure 1 for dimensions and nomenclature.

As a first selection, the liquid level hL should be set to 2/3 of the vessel internal
diameter D . A first estimate of D can be made from the desired residence time θ A and
liquid throughput Q A . This first estimate should be made in anticipation of obtaining an
acceptable slenderness ratio for the vessel in the final design.

(3) Crest height, velocity head, friction loss and oil layer thickness

The crest height and velocity of the liquid flow over the overflow weir can be calculated
using the following implicit equations given in Daugherty & Franzini (11):

209
  2  Vo2  2 
3 3
2
2 V
Q A = C D 2 g B  H + o
 −   
3  2 g   2 g  
 
1 H
C D = 0.605 + + 0.08
H hA

Where:

B is the width of the overflow weir; and in the expression for the weir discharge
coefficient C D , H and h A are in mm.

Also given in Daugherty & Franzini(11) is a simplified explicit, less accurate expression
relating Q A and H . The explicit expression avoids iteration and has been used in the
procedure developed in the past for manual calculations. However, nowadays the
iteration required by the implicit expression can be implemented easily with a
spreadsheet.

To calculate the flow area of the crest of height H and hence velocity of approach Vo (the
horizontal velocity of the liquid flowing in the crest is known as the velocity of
approach), the following relation in circular geometry may be helpful:

b−r 
A = (b − r ) r 2 − (b − r ) + r 2 sin −1 
2
−
r  r 
 −1  a − r  
(a − r ) r − (a − r ) + r sin 
2 2 2
a b 
  r 

Where:
A is the area of the strip of the circle (of radius r ) defined by a and b ; and
a = hA , b = hA + H .

The head loss incurred by the underflow and upflow is:

VU2 h V2
hF = + 4 f A UP
2g DE 2 g

The first term on the right represents the loss incurred in flowing horizontally through the
underflow opening and making the 90 ° turn upward; and VU is the underflow velocity.
The second term is the friction loss incurred in flowing upward through the upflow
channel. Without significant loss of accuracy, the flow area in the upflow channel is
taken to be WUP B . The upflow velocity VUP , the hydraulic (or equivalent) diameter DE
and the Fanning friction factor f are all based on this flow area WUP B . The hydraulic

210
diameter of a flow area is defined as: (4 x flow area) / (wetted perimeter of flow area).
For D E we have:

4 × WUP B
DE =
2WUP + 2 B

For Fanning friction factor, there are excellent explicit empirical formulas are available:
Goudar(13) and Coker(14).

The oil layer thickness y is given by:


ρA   Vo2 
y= (hL − h A ) − H −  + hF 
ρ A − ρ o   2g 

(4) Acceptable crest height and oil layer thickness

The acceptable oil layer thickness y is 100 mm; that of crest height H is 25 to 75 mm. If
acceptable values cannot be obtained by adjusting h A , revise D , hL and repeat.

(5) Acceptable friction loss

The friction loss hF should be kept below 3 mm. This can be easily achieved with
reasonable dimensions of the underflow weir clearance and the upflow channel width. It
is reasonable to take the underflow clearance and upflow channel width to be 12 to 15 %
of the vessel diameter D.

(6) Liquid droplet terminal velocity

The terminal velocity Vt of the liquid droplet falling through the gas phase (Figure 1) is
obtained by iteration of the following expressions relating Vt , C D and Re .

For a liquid droplet falling down in a gas phase:

Vt 2
FD = C D Aρ G
2
πd m3
FB = (ρ L − ρ G )g
6
Equating FD and FB :

211
 ρ − ρ G  4d m
Vt =  L  g FD
 ρ G  3C D
24 3
CD = + 0.5 + 0.34 Vt
Re Re
ρ Vd FB

Re = G t m Liquid droplet
µG falling in gas phase

Where:
FD : drag force on the falling droplet
FB : weight of droplet in vapour space
C D : drag coefficient
A : cross section of droplet perpendicular to the direction of fall, πD 2 4
Re : Reynolds number of falling droplet

(7) Gas and liquid constraints

For a selected diameter D , we define the following ratios:

hL
α=
D
2 AG
A 1  2h  2  2h  h  h  y
β = L = cos −1 1 − L  − 1 − L  L −  L 
AC π  D  π D  D D Ao
D

2
AW 1  2h  2  2hW  hW  hW  hL hw AW
βW = = cos −1 1 − W  − 1 −  − 
AC π  D  π D  D  D
A L = Ao + AW

Where:
AG : flow area of gas phase
AL : flow area of liquid phase (oil + aqueous)
Ao : flow area of oil phase
AW : flow area of aqueous phase
AC : total circular area, πD 2 4

From these relations in circular geometry, we proceed as follows to calculate the gas
constraint L*G D and liquid constraint L*W D 2 ; L*G and L*W are the minimum path lengths
required to meet the residence times of the gas and the aqueous phases, respectively; one
of them is governing. For a rich amine flash drum, where only the dissolved gas is
flashed off, L*W is invariably governing.

212
Gas constraint

Equating the time of droplet fall with the gas phase residence time:

L*G D(1 − α )
=
VG Vt

Vt is the terminal velocity of the droplet falling in the gas phase and VG is the horizontal
gas velocity in the vessel given by:

4QG
VG =
(1 − β )πD 2
We obtain:

L*G D =
(1 − α ) 4QG
(1 − β ) πVt
Liquid constraint

From the desired aqueous phase residence time θ W we have the relation:
QW θ W πD 2
AW = = β W
L*W 4
From it we obtain:
4QW θ W
L*W D 2 =
βW π

From the selected diameter D , L*G and L*W are obtained, one of which will be governing.

(8) Surge compartment length

From the desired surge time θ S to be provided for the liquid held in the surge
compartment, the length of the required for the surge compartment length LS can be
calculated. The following relation is useful to make this calculation.

The volume V in the compartment of a horizontal vessel with a 2:1 SE head (2 to 1


spherical elliptical head of a vessel) is given by:

213
L
π D h D2  2h  D  2h   2:1 SE head
V = h 2  −  + LS  cos −1 1 −  − 1 −  Dh − h 2 
4  2 3  4  D 2 D  D
V h

The first term on the right is the volume held in the head; the second term is the volume
held in the cylindrical part of the compartment; and h is the liquid level.

(9) Overall vessel length

In a skimmer application such as amine flash drum, it is usually the requirement of the
aqueous phase residence time θ W that governs the length of the residence compartment,
i.e. L*W > L*G . The residence compartment length together with the upflow channel width
WUP and the surge compartment length LS give the overall tangent-to-tangent length LTT of
the vessel.

(10) Slenderness ratio

A slenderness ratio LTT D should be obtained for a horizontal vessel in this application in
consideration of the requirements and constraints of footprint, fabrication and
transportation. Usually when these requirements and constraints are met, the slenderness
ratio turns out to be in the range of 3 to 5.

(11) Rating at 120 % throughput

With the final selected dimensions, the design should be rated at 120 % the aqueous
phase throughput. As throughput increases, the oil layer thickness y diminishes. At 120 %
a minimum oil layer thickness of 25 mm should be maintained. If this minimum cannot
be achieved, repeat the iteration.

214
V.4 Calculation Flowchart

Given: (1)
Start Design parameters

Select: (2)
Vessel ID D
End
Liquid height hL

Yes No
Select:
y ≥ 25 mm? Underflow clearance hU
Upflow channel w idth W UP

Select:
Rate at 120 % Aqueous overflow
throughput (11) w eir height hA
Yes No
Calculate: (3)
Slenderness Crest height H
LT T /D OK? Velocity head V o2/2g
(10) Friction loss hF
Oil layer thickness y

Select: (9)
Overall vessel length LT T Are y, H (4) No
acceptable?

Yes

Calculate: (8) Is hF (5) No


Surge compartment
acceptable?
Length LS

Yes
Calculate:(7) Calculate: (6)
Terminal velocity of liquid
Gas constraint L* G
droplets falling through
*
Liquid constraint L W
gas phase Vt

215
Appendix VI:

Pinch Temperature and Exchanger Duty Shortfall

Presented in this appendix are the calculations of the pinch temperature and heat duty
shortfall of the cocurrent arrangement discussed in Section 5.7.

Premises:
The heat transfer area of the heat exchanger is such that the process targets of heat duty
and temperatures are met in the countercurrent arrangement. The exchanger in the
countercurrent arrangement is identical to the one in the cocurrent arrangement. The mass
of flow rates and inlet conditions of the hot and cold fluids are identical in both
arrangements. Both fluids remain single-phase and their heat capacities are treated as
constant throughout the heat exchange process. Also, implicit in the following relations
and equations is that the heat transfer coefficients remain constant.

T1 = 75 oC

o
LMTD = 16.6 C
Temperture

Hot inlet, T1

75 C
o
o
T2 = 33 oC, target met
t 2 = 45 C,
Cold inlet, t1 target met
o
25 C
Intended duty achieved t1 = 25 oC
Cold outlet, t 2 Hot outlet, T2
o o
45 C 33 C
Heat transacted
Countercurrent arrangement
Countercurrent temperature profile and LMTD

o
T1 = 75 C o
LMTD = 13.1 C
Hot inlet, T1
o
o T2C= 41.9 C, target missed
Temperture

75 C
Cold outlet, t C T pinch = 41.1 oC
o 2
40.7 C
Cold inlet, t1 Hot outlet, T2C t 2C o
= 40.7 C, target missed
25 C
o o
41.9 C
t1 = 25 o
C
Only 79% of
intended duty
Cocurrent arrangement
Heat transacted
Cocurrent temperature profile and LMTD
Figure 5.13 :
Cocurrent pinch causing a shortfall in duty

216
(a) Pinch Temperature

A temperature pinch exists with the cocurrent arrangement when it is intended that hot
fluid outlet temperature T 2 be lower than the cold fluid outlet temperature t 2 .

We have by the heat flow equations:

(T − T )(m C )
1 pinch p hot =Q
(t − T )(m C )
1 pinch p cold =Q

Or:

T1 − T pinch (m C )
p cold
= (1)
T pinch − t1 (m C )
p hot

The quantities m , C P are not given explicitly in this example, but are implicit in the
temperature profile constructed from the inlet and outlet temperatures as shown in the
sketch above (Figure 5.13). Implied in the temperature profiles is that there is no phase
chase change in either the hot or cold fluid; and that the specific heat of either the hot or
cold fluid remains constant. We cannot therefore solve (1) explicitly to obtain the pinch
temperature T pinch . We can, however, find T pinch by the following iterative procedure:

1. Guess a hot fluid outlet temperature T2*


2. Calculate cold fluid out temperature t 2* based on T2* as:
t −t
(
t 2* = t1 + T1 − T2* 2 1
T1 − T2
)
3. If t 2* = T2* , then T pinch = t 2* = T2* ; otherwise repeat Steps 1 to 2.
(For the example in Figure 5.13, convergence yields T pinch = 41.1 °C.)

(b) Heat duty shortfall

To obtain the duty shortfall we need first to calculate the hot and cold outlet temperatures
T2C and t 2C for the cocurrent arrangement. We do it by iteration as demonstrated below:

We have by heat balance:

Qcount = UA × LMTDcount = (m C p )hot (T1 − T2 ) ;


where T1 − T2 is the temperature drop (on target) of hot fluid of the
counter-current arrangement

217
(
Qco = UA × LMTDco = (m C p )hot T1 − T2c ; )
where T1 − T2c is the temperature drop (with shortfall) of hot fluid in the cocurrent
arrangement

Dividing one of the preceding two equations by the other, we have:

T1 − T2c
LMTDco = LMTDcount (2)
T1 − T2

We can find the hot fluid outlet temperature T2c and cold fluid outlet temperature t 2c of the
cocurrent arrangement by the following iterative procedure:

1. Guess hot fluid outlet temperature T2c

2. Calculate LMTDco using (2) as:

T1 − T2c
LMTDco = LMTDcount ; and, by its definition, we have:
T1 − T2
(T − t ) − (T2 − t1 )
LMTDcount = 1 2
T −t
ln 1 2
T2 − t1
3. Calculate cold fluid outlet temperature t 2c as:
t −t
( )
t 2c = t1 + T1 − T2c 2 1
T1 − T2
4. Calculate LMTDco by its definition as:

(T1 − t1 ) − (T2c − t 2c )
LMTDco =
T1 − t1
ln
T2c − t 2c
5. If LMTDco by Step 2 = LMTDco by Step 4, then T2c and t 2c are the converged hot
and cold fluid outlet temperatures, respectively, for the cocurrent arrangement;
proceed to Step 6; otherwise repeat Step 1 to 4 to reach convergence.
[For the example in Figure 5.13, convergence yields T2c = 41.9 °C and t 2c = 40.7
°C]

218
6. Calculate the heat duty shortfall as:

Qco T1 − T2c
1− = 1−
Qcount T1 − T2

(For the example in Figure 5.13, the duty shortfall is 21 %.)

Nomenclature:

Cp : heat capacity
LMTD : log mean temperature
m : mass flow rate
Q : heat duty
T : temperature of hot fluid
t : temperature of cold fluid
∆ : difference in

Subscripts
co : cocurrent
cold : cold
count : countercurrent
hot : hot fluid
pinch : at (temperature) pinch
1 : inlet
2 : outlet

Superscripts
* : guessed value
c : cocurrent

219
Appendix VII:

LMTD Correction

The LMTD correction is illustrated here for the example in Section 5.7 Figure 5.15. Both
an empirical formula and charts are used to find the LMTD correction factor.

(a) Two-pass tube bundle in one TEMA E shell

Using empirical formula


Hot fluid IN temperature T1 75.0 t 2 − t1
P=
Hot fluid OUT temperature T2 33.0 T1 − t 1
Cold fluid IN temperature t1 25.0
Cold fluid OUT temperature t2 45.0 T1 − T2
R =
t 2 − t1
Number of shells in series N 1
P 0.40 1/ N
 RP - 1 
R 2.10 1−  
 P -1 
Px 0.40
Px = 1/ N
 RP - 1 
LMTD correction factor F 0.010 R − 
 P -1 
LMTD uncorrected LMTD 16.64
LMTD Corrected CorrMTD 0.17
F has fallen off the cliff; increase number of shells !

F=
(R 2
+1 ) 0.5

ln
1 − Px
÷ ln
(
2 / Px − 1 − R + R 2 + 1 )
0.5

R −1 1 − RPx (
2 / Px − 1 − R − R 2 + 1
0.5
)
Using chart

(b) Two TEMA


F, LMTD correction factor

In these charts , HOT is shown as shellside fluid and


COLD as tubeside. The LMTD correction is valid as well
with hot fluid in tube side and cold fluid in shell side.

P, temperature efficiency

P = 0.40
T1 R = 2.10
t2 F = 0.01, fallen off the cliff
1 shell pass per shell (TEMA E)
& off-scale
t1 2 tube passes per shell
T2

220
(b) Two TEMA E shells in series

Using empirical formula

Hot fluid IN temperature T1 75.0


t 2 − t1
Hot fluid OUT temperature T2 33.0 P=
T1 − t 1
Cold fluid IN temperature t1 25.0
Cold fluid OUT temperature t2 45.0 T1 − T2
R =
t 2 − t1
Number of shells in series N 2
P 0.40 1/ N
 RP - 1 
R 2.10 1−  
Px 0.31  P -1 
Px = 1/ N
LMTD correction F 0.853  RP - 1 
R − 
LMTD uncorrected LMTD 16.64  P -1 
LMTD Corrected CorrMTD 14.21
F is acceptable
F=
(R 2
+1 ) 0.5

ln
1 − Px
÷ ln
(
2 / Px − 1 − R + R 2 + 1 )
0.5

R −1 1 − RPx (
2 / Px − 1 − R − R 2 + 1
0.5
)

Using chart

P = 0.40
R = 2.10
F, LMTD correction factor

F = 0.853

In these charts , HOT is shown as shellside fluid and


COLD as tubeside. The LMTD correction is valid as well
with hot fluid in tube side and cold fluid in shell side.

P, temperature efficiency
T1
t2

2 shells in series (1 shell pass per shell, TEMA E)


2 tube passes per shell
t1
T2

221
Appendix VIII:

Thermodynamic Relations (relating state functions to measurable PVT and calorific


data C P and C V )

We need to express the state functions U, S, H, G, A in terms of integrals that can be


evaluated using equations of states constructed from measured PVT and specific heat
data. The expressions for these state functions can be derived using the differential forms
of the first law together with a family of relations known as the Maxwell relations.

First Law in differential forms (see Chapter 7):

dU = TdS − PdV (1)

dH = TdS + VdP (2)

Maxwell relations (a partial list only to meet the need here):

 ∂S   ∂V 
  = −  (3)
 ∂P  T  ∂T  P

 ∂S   ∂P 
  =  (4)
 ∂V  T  ∂T V

Specific heats

The specific heats at constant pressure C P and at constant volume C V are by definition:

 ∂H 
CP =   (5)
 ∂T  P

 ∂U 
CV =   (6)
 ∂T V

We illustrate here the derivation of the expressions for U(T,V) and U(T, P) and list the
rest, which can be derived in a similar manner.

U(T, V)

Considering U as a function of T, V, i.e. U = U(T, V), we can write from the first principle
of calculus:

 ∂U   ∂U 
dU (T , V ) =   dT +   dV
 ∂T V  ∂V  T

222
 ∂U 
= CV dT +   dV , using (6) (7)
 ∂V  T

The first law written in differential form (1) can be re-cast as:

 ∂U   ∂S 
  = T  −P
 ∂V  T  ∂V  T
 ∂P 
= T  − P , using Maxwell relation (4) (8)
 ∂T V

Combining (7) and (8) and integrating, we have:

T V
  ∂P  
U (T , V ) − U (T0 , V0 ) = ∫ CV dT + ∫ T   − P  dV (9)
T0 V0  
∂T V 

We have used (P 0 ,V 0 ,T 0 ) as the reference condition.

U(T, P)

With U expressed as a function of T and P, i.e. U = U(T, P), we have again by the first
principle of calculus:

 ∂U   ∂U 
dU (T , P ) =   dT +   dP (10)
 ∂T  P  ∂P  T

From the definition H = U + PV , we have:

 ∂U   ∂H   ∂V 
  =  − P 
 ∂T  P  ∂T  P  ∂T  P
 ∂V 
= C P − P  , using (5); and (11)
 ∂T  P

 ∂U   ∂H   ∂V 
  =  − P  −V (12)
 ∂P  T  ∂P  T  ∂P  T

From (2), we have:

 ∂H   ∂S 
  = T  +V
 ∂P  T  ∂P  T

223
 ∂V 
= −T   + V , using Maxwell relation (3) (13)
 ∂T  P
Combing (12) and (13), we have:

 ∂U   ∂V   ∂V 
  = −T   − P  (14)
 ∂P  T  ∂T  P  ∂P  T

Substituting (11) and (14) into (10) and integrating we have:

T P
  ∂V     ∂V   ∂V  
U (T , P ) − U (T0 , P0 ) = ∫ C P − P   dT − ∫ T   + P  dP (15)
T0   ∂T  P  P0  
∂T  P  ∂P  T 

List of Thermodynamic States in Terms of PVT, C P and C V

For convenience, we collect the results (9) and (15) and list the rest here:

T V
  ∂P  
U (T , V ) = ∫ CV dT + ∫ T   − P  dV (9)
T0 V0  
∂T V 
T P
  ∂V     ∂V   ∂V  
U (T , P ) = ∫ C P − P   dT − ∫ T   + P  dP (15)
T0   ∂T  P  P0  
∂T  P  ∂P  T 
T V
CV  ∂P 
S (T , V ) = ∫ dT + ∫   dV (16)
T0
T V0 
∂T V
T P
CP  ∂V 
S (T , P ) = ∫T T dT − ∫P  ∂T  P dP (17)
0 0

T V
  ∂P     ∂P   ∂P  
H (T ,V ) − H (T0 ,V0 ) = ∫ CV + V   dT + ∫ T   +V   dV (18)
T0   ∂T V  V0  
∂T V  ∂V  T 
T P
  ∂V  
H (T , P) − H (T0 , P0 ) = ∫ C P dt + ∫ V − T    dP (19)
T0 P0   ∂T  P 

Note that all the integrands in the list above are obtainable from measured PVT data and
measured specific heats.

The state functions G and A can be obtained by their definitions: G = H – TS and


A = U – TS.

Note that if we set U(T 0 ,V 0 ) =U(T 0 , P 0 ) to an arbitrary value, H(T 0 ,V 0 ) =H(T 0 , P 0 )


cannot have an arbitrary value since, by definition H = U + PV, H(T 0 ,V 0 ) = U(T 0 ,V 0 )
+ P 0 V 0 ; and, H(T 0 ,P 0 ) = U(T 0 ,P 0 ) + P 0 V 0 .

224
Appendix IX:

Batch Boiler:
What is the heat input rate q1 required to make steam at the rate of m 2 ?

Pressure P = 475.8 kPa (a)


Surroundings m 2 , H 2
w S =0
Temperature T = 150 o C
2
System Vapour Liquid Point 2
q 2 = 0 Enthalpy, kJ/kg: HV HL H 2 (= H V )
Sat'd steam
m 1 = 0
Internal energy, kJ/kg: U
V UL U 2 (= U V )
q1
V2 (= VV )
3
Sat'd w ater
1
Specific volume, m / kg: VV VL
H1
Hold-up, kg: mV mL
Steam production, kg/s : m 2 = 20
Figure 1:
Making steam in batch boiler Total hold-up in system, kg: mC = mV + m L

First law (energy balance):

 u2   u2  d
m 2 U 2 + P2V2 + 2 + z 2 g  − m 1 (U 1 + P1V 1 ) + 1 + z1 g  = q − w S − (mC U C )
 2   2  dt

Set
Since: m1 = 0; w S = 0; u 22 2 → 0, z 2 g → 0 , we have:
d
q = m 2 H 2 + (mCU C ) ; or
dt
d
q1 = m 2 H 2 + (mCU C ) (1)
dt
Since q = q1 − q 2 and q 2 = 0 (no heat leaving the system).
 

Considering a change in liquid hold-up δ m L (kg) in the batch boiler and a concomitant
change in vapour hold-up δmV (kg) occurring in time interval δ t , we have:

Volume change in liquid hold-up: VL δ m L

Volume change in vapour hold-up: VV δ mV

VV δ mV = −V Lδ m L (constant total system volume); or:

225
dmV V dm L
=− L (2)
dt VV dt

Also:
mC = mV + m L

dmC dm dm L
m 2 = − =− V − (3)
dt dt dt

Combining (2) and (3), to eliminate dmV dt :

dm L m 2
= (4)
dt (VL VV ) − 1
(A sanity check on (4): as VL VV is much smaller than 1, m 2 ≈ − dm L dt ; or, rate of steam
leaving is, to a good approximation, equal to mass depletion rate of liquid hold-up.)

Since mCU C = mV U V + m LU L , we have:

d
(mCU C ) = U V dmV + U L dm L (5)
dt dt dt

Substituting (2) and (4) into (5), we have the rate of internal energy accumulation as:

d
(mCU C ) = m 2 U L − U V VL VV (6)
dt (VL VV ) − 1
Combining the specialized expression of the first law (1) and (6), we get:

U L − U V VL VV
q1 = m 2 H 2 + m 2 (7)
(VL VV ) − 1
From (7) we can determine the heat input rate q1 required from the given values of H 2 ,
U L , U V , V L , V V and m 2 .

226
References

1. Sloley A. W., “Properly Design Thermosiphon Reboilers”, Chemical Engineering


Progress, March 1997

2. Sloley A. W., “Steady Under Pressure”, AIChE Spring Meeting


6-9 March 2000

3. Arnold & Stewart, “Design of Oil-handling Systems and Facilities”, Gulf


Publishing Company (Three-phase Oil and Gas Separators, pp126+)

4. Buzzetta P. et al, “Control optimization saves money in distillation”,


Hydrocarbon Processing, June 2007

5. Chin T. G., “Guide to Distillation Pressure Control Methods”, Hydrocarbon


Processing, October 1979

6. Kister H., “Distillation Design”, McGraw-Hill Professional, 1992, ISBN:


0070349096; (Reboiler, Condenser and Pressure Controls) pp530+

7. Friedman, Y. Zak, “APC design for minimum maintenance”, Hydrocarbon


Processing, July 2009.

8. Callen H. B., “Thermodynamics”, John Wiley & Sons, Inc. (1960)

9. Boas, Mary L., “Mathematical Methods in the Physical Sciences”, John Wiley &
Sons, Inc. (1983)

10. Einstein A., Szilard L., “Einstein Refrigerator”, US Patent 1781541 Nov 11, 1930

11. Daugherty and Franzini, “Fluid Mechanics with Engineering Applications (6th
ed.), McGraw-Hill Book Company, Kogakusha Company Ltd (1965)

12. Giles, R.V., Evett, J. B., Liu, C., “Fluid Mechanics and Hydraulics” ( 3rd ed.),
Schaum’s Outline Series, McGraw Hill (1994)

13. Goudar, C. T., Sonnad, J. R., “Explicit Friction Factor Correlation for Turbulent
Flow in Rough Pipe”, Hydrocarbon Processing, April 2007, pp 103 –

14. Coker, A. K., “Understanding Two-Phase Flow in Process Piping”, Chemical


Engineering Progress, Nov 1990, pp 60 -

227

You might also like