Вы находитесь на странице: 1из 52

Steam Distribution

Efficient distribution gets clean dry steam to apparatus at the right pressure. Pipe sizing, essential
drainage techniques, pipe support and expansion, air venting, and heat transfer calculations are included
to help the system designer and practitioner.

1. Introduction to Steam Distribution


An efficient steam distribution system is essential if steam of the right quality and pressure is to
be supplied, in the right quantity, to the steam using equipment. This tutorial looks at a typical
circuit.

2. Pipes and Pipe Sizing


Pipe sizing is a crucial aspect of steam system design. This tutorial offers detailed advice on
standards, schedules, materials and sizing for various saturated and superheated steam duties.

3. Steam Mains and Drainage


Issues surrounding the structure, layout and operation of a steam distribution system, including
condensate drain points and branch lines, the avoidance of waterhammer and the use of
separators and strainers for steam conditioning.

4. Pipe Expansion and Support


Any steam system must be fully supported, able to expand during operation and sufficiently
flexible to allow movement as a result. This tutorial includes advice on different methods and full
calculations.

5. Air Venting, Heat Losses and a Summary of Various Pipe


Related Standards
The venting of air and other incondensable gases from steam systems, and the provision of
adequate insulation, are vital to ensure steam plant efficiency, safety and performance.
Introduction to Steam Distribution
The steam distribution system is the essential link between the steam generator and the steam user.

This Tutorial will look at methods of carrying steam from a central source to the point of use. The central
source might be a boiler house or the discharge from a co-generation plant. The boilers may burn primary
fuel, or be waste heat boilers using exhaust gases from high temperature processes, engines or even
incinerators. Whatever the source, an efficient steam distribution system is essential if steam of the right
quality and pressure is to be supplied, in the right quantity, to the steam using equipment. Installation and
maintenance of the steam system are important issues, and must be considered at the design stage.

Top

Steam system basics


From the outset, an understanding of the basic steam circuit, or 'steam and condensate loop' is required -
see Figure 10.1.1. As steam condenses in a process, flow is induced in the supply pipe. Condensate has a
very small volume compared to the steam, and this causes a pressure drop, which causes the steam to
flow through the pipes.

Fig. 10.1.1 A typical basic steam circuit


The steam generated in the boiler must be conveyed through pipework to the point where its heat energy
is required. Initially there will be one or more main pipes, or 'steam mains', which carry steam from the
boiler in the general direction of the steam using plant. Smaller branch pipes can then carry the steam to
the individual pieces of equipment.

When the boiler main isolating valve (commonly called the 'crown' valve) is opened, steam immediately
passes from the boiler into and along the steam mains to the points at lower pressure. The pipework is
initially cooler than the steam, so heat is transferred from the steam to the pipe. The air surrounding the
pipes is also cooler than the steam, so the pipework will begin to transfer heat to the air.

Steam on contact with the cooler pipes will begin to condense immediately. On start-up of the system, the
condensing rate will be at its maximum, as this is the time where there is maximum temperature difference
between the steam and the pipework. This condensing rate is commonly called the 'starting load'. Once
the pipework has warmed up, the temperature difference between the steam and pipework is minimal, but
some condensation will occur as the pipework still continues to transfer heat to the surrounding air. This
condensing rate is commonly called the 'running load'.

The resulting condensation (condensate) falls to the bottom of the pipe and is carried along by the steam
flow and assisted by gravity, due to the gradient in the steam main that should be arranged to fall in the
direction of steam flow. The condensate will then have to be drained from various strategic points in the
steam main.

When the valve on the steam pipe serving an item of steam using plant is opened, steam flowing from the
distribution system enters the plant and again comes into contact with cooler surfaces. The steam then
transfers its energy in warming up the equipment and product (starting load), and, when up to temperature,
continues to transfer heat to the process (running load).

There is now a continuous supply of steam from the boiler to satisfy the connected load and to maintain
this supply more steam must be generated. In order to do this, more water (and fuel to heat this water) is
supplied to the boiler to make up for that water which has previously been evaporated into steam.

The condensate formed in both the steam distribution pipework and in the process equipment is a
convenient supply of useable hot boiler feedwater. Although it is important to remove this condensate from
the steam space, it is a valuable commodity and should not be allowed to run to waste. Returning all
condensate to the boiler feedtank closes the basic steam loop, and should be practised wherever practical.
The return of condensate to the boiler is discussed further in Block 13, 'Condensate Removal', and Block
14,'Condensate Management'.

The working pressure


The distribution pressure of steam is influenced by a number of factors, but is limited by:

 The maximum safe working pressure of the boiler.


 The minimum pressure required at the plant.
As steam passes through the distribution pipework, it will inevitably lose pressure due to:

 Frictional resistance within the pipework (detailed in Tutorial 10.2).


 Condensation within the pipework as heat is transferred to the environment.
Therefore allowance should be made for this pressure loss when deciding upon the initial distribution
pressure.

A kilogram of steam at a higher pressure occupies less volume than at a lower pressure. It follows that, if
steam is generated in the boiler at a high pressure and also distributed at a high pressure, the size of the
distribution mains will be smaller than that for a low-pressure system for the same heat load. Figure 10.1.2
illustrates this point.

Fig. 10.1.2 Dry saturated steam -


pressure/specific volume relationship
Generating and distributing steam at higher pressure offers three important advantages:

 The thermal storage capacity of the boiler is increased, helping it to cope more efficiently
with fluctuating loads, minimising the risk of producing wet and dirty steam.
 Smaller bore steam mains are required, resulting in lower capital cost, for materials such
as pipes, flanges, supports, insulation and labour.
 Smaller bore steam mains cost less to insulate.
Having distributed at a high pressure, it will be necessary to reduce the steam pressure to each zone or
point of use in the system in order to correspond with the maximum pressure required by the application.
Local pressure reduction to suit individual plant will also result in drier steam at the point of use. (Tutorial
2.3 provides an explanation of this).

Note: It is sometimes thought that running a steam boiler at a lower pressure than its rated pressure will
save fuel. This logic is based on more fuel being needed to raise steam to a higher pressure.

Whilst there is an element of truth in this logic, it should be remembered that it is the connected load, and
not the boiler output, which determines the rate at which energy is used. The same amount of energy is
used by the load whether the boiler raises steam at 4 bar g, 10 bar g or 100 bar g. Standing losses, flue
losses, and running losses are increased by operating at higher pressures, but these losses are reduced
by insulation and proper condensate return systems. These losses are marginal when compared to the
benefits of distributing steam at high pressure.

Pressure reduction
The common method for reducing pressure at the point where steam is to be used is to use a pressure
reducing valve, similar to the one shown in the pressure reducing station Figure 10.1.3.

F
ig. 10.1.3 Typical pressure reducing valve station
A separator is installed upstream of the reducing valve to remove entrained water from incoming wet
steam, thereby ensuring high quality steam to pass through the reducing valve. This is discussed in more
detail in Tutorial 9.3 and Tutorial 12.5.

Plant downstream of the pressure reducing valve is protected by a safety valve. If the pressure reducing
valve fails, the downstream pressure may rise above the maximum allowable working pressure of the
steam using equipment. This, in turn, may permanently damage the equipment, and, more importantly,
constitute a danger to personnel.

With a safety valve fitted, any excess pressure is vented through the valve, and will prevent this from
happening (safety valves are discussed in Block 9).

Other components included in the pressure reducing valve station are:

 The primary isolating valve - To shut the system down for maintenance.
 The primary pressure gauge - To monitor the integrity of supply.
 The strainer - To keep the system clean.
 The secondary pressure gauge - To set and monitor the downstream pressure.
 The secondary isolating valve - To assist in setting the downstream pressure on no-
load conditions.
Pipes and Pipe Sizing
Pipe sizing is a crucial aspect of steam system design. This tutorial offers detailed advice on
standards, schedules, materials and sizing for various saturated and superheated steam duties.

Use the quick links below to take you to the main sections of this tutorial:

 Pipe material
 Pipeline sizing
 Steam
 Summary
 Appendix
 The printable version of this page has now been replaced by
The Steam and Condensate Loop Book
 Try answering the Questions for this tutorial
 View the complete collection of Steam Engineering Tutorials
 Contact Us

Standards and wall thickness


There are a number of piping standards in existence around the world, but arguably the most global are
those derived by the American Petroleum Institute (API), where pipes are categorised in schedule
numbers.

These schedule numbers bear a relation to the pressure rating of the piping. There are eleven Schedules
ranging from the lowest at 5 through 10, 20, 30, 40, 60, 80, 100, 120, 140 to schedule No. 160. For
nominal size piping 150 mm and smaller, Schedule 40 (sometimes called 'standard weight') is the lightest
that would be specified for steam applications.

Regardless of schedule number, pipes of a particular size all have the same outside diameter (not
withstanding manufacturing tolerances). As the schedule number increases, the wall thickness increases,
and the actual bore is reduced. For example:

 A 100 mm Schedule 40 pipe has an outside diameter of 114.30 mm, a wall thickness of
6.02 mm, giving a bore of 102.26 mm.
 A 100 mm Schedule 80 pipe has an outside diameter of 114.30 mm, a wall thickness of
8.56 mm, giving a bore of 97.18 mm.
Only Schedules 40 and 80 cover the full range from 15 mm up to 600 mm nominal sizes and are the most
commonly used schedule for steam pipe installations.

This Tutorial considers Schedule 40 pipework as covered in BS 1600.

Tables of schedule numbers can be obtained from BS 1600 which are used as a reference for the nominal
pipe size and wall thickness in millimetres. Table 10.2.1 compares the actual bore sizes of different sized
pipes, for different schedule numbers.

In mainland Europe, pipe is manufactured to DIN standards, and DIN 2448 pipe is included in Table
10.2.1.
Table 10.2.1 Comparison of pipe standards and actual bore diameters.
In the United Kingdom, piping to EN 10255, (steel tubes and tubulars suitable for screwing to BS 21
threads) is also used in applications where the pipe is screwed rather than flanged. They are commonly
referred to as 'Blue Band' and 'Red Band'; this being due to their banded identification marks. The different
colours refer to particular grades of pipe:

 Red Band, being heavy grade, is commonly used for steam pipe applications.
 Blue Band, being medium grade, is commonly used for air distribution systems, although
it is sometimes used for low-pressure steam systems.
The coloured bands are 50 mm wide, and their positions on the pipe denote its length. Pipes less than 4
metres in length only have a coloured band at one end, while pipes of 4 to 7 metres in length have a
coloured band at either end.

Fig. 10.2.1 Red band, branded pipe, - heavy grade

Fig. 10.2.2 Blue band, branded pipe, - medium grade, between 4-7
metres in length
Top

Pipe material
Pipes for steam systems are commonly manufactured from carbon steel to ANSI B 16.9 A106. The same
material may be used for condensate lines, although copper tubing is preferred in some industries.

For high temperature superheated steam mains, additional alloying elements, such as chromium and
molybdenum, are included to improve tensile strength and creep resistance at high temperatures.

Typically, pipes are supplied in 6 metre lengths.

Top

Pipeline sizing
The objective of the steam distribution system is to supply steam at the correct pressure to the point of
use. It follows, therefore, that pressure drop through the distribution system is an important feature.

Liquids
Bernoulli's Theorem (Daniel Bernoulli 1700 - 1782) is discussed in Block 4 - Flowmetering. D'Arcy (D'Arcy
Thompson 1860 - 1948) added that for fluid flow to occur, there must be more energy at Point 1 than Point
2 (see Figure 10.2.3). The difference in energy is used to overcome frictional resistance between the pipe
and the flowing fluid.
F
ig. 10.2.3 Friction in pipes
Bernoulli relates changes in the total energy of a flowing fluid to energy dissipation expressed
either in terms of a head loss hf (m) or specific energy loss g hf (J/kg). This, in itself, is not very
useful without being able to predict the pressure losses that will occur in particular
circumstances.

Here, one of the most important mechanisms of energy dissipation within a flowing fluid is
introduced, that is, the loss in total mechanical energy due to friction at the wall of a uniform
pipe carrying a steady flow of fluid.

The loss in the total energy of fluid flowing through a circular pipe must depend on:
L = The length of the pipe (m)
D = The pipe diameter (m)
u = The mean velocity of the fluid flow (m/s)
 = The dynamic viscosity of the fluid (kg/m s=Pa s)
ρ = The fluid density (kg/m3)
ks = The roughness of the pipe wall* (m)

*Since the energy dissipation is associated with shear stress at the pipe wall, the nature of the wall surface
will be influential, as a smooth surface will interact with the fluid in a different way than a rough surface.

All these variables are brought together in the D'Arcy-Weisbach equation (often referred to as the D'Arcy
equation), and shown as Equation 10.2.1. This equation also introduces a dimensionless term referred to
as the friction factor, which relates the absolute pipe roughness to the density, velocity and viscosity of the
fluid and the pipe diameter.

The term that relates fluid density, velocity and viscosity and the pipe diameter is called the Reynolds
number, named after Osborne Reynolds (1842-1912, of Owens College, Manchester, United Kingdom),
who pioneered this technical approach to energy losses in flowing fluids circa 1883.

The D'Arcy equation (Equation 10.2.1):

Equation 10.2.1
Where:
hf = Head loss to friction (m)
f = Friction factor (dimensionless)
L = Length
u = Flow velocity (m/s)
g = Gravitational constant (9.81 m/s2)
D = Pipe diameter (m)
Interesting point
Readers in some parts of the world may recognise the D'Arcy equation in a slightly different form, as
shown in Equation 10.2.2. Equation 10.2.2 is similar to Equation 10.2.1 but does not contain the constant
4.

Equation 10.2.2
The reason for the difference is the type of friction factor used. It is essential that the right version of the
D'Arcy equation be used with the selected friction factor. Matching the wrong equation to the wrong friction
factor will result in a 400% error and it is therefore important that the correct combination of equation and
friction factor is utilised. Many textbooks simply do not indicate which friction factors are defined, and a
judgement must sometimes be based on the magnitudes quoted.

Equation 10.2.2 tends to be used by those who traditionally work in Imperial units, and still tends to be
used by practitioners in the United States and Pacific rim regions even when metric pipe sizes are quoted.
Equation 10.2.1 tends to be used by those who traditionally work in SI units and tends more to be used by
European practitioners. For the same Reynolds number and relative roughness, the 'Imperial based friction
factor' will be exactly four times larger than the 'SI based friction factor'.

Friction factors can be determined either from a Moody chart or, for turbulent flows, can be calculated from
Equation 10.2.3, a development of the Colebrook - White formula.

Equation 10.2.3
Where:
f = Friction factor (Relates to the SI Moody chart)
ks = Absolute pipe roughness (m)
=
D Pipe bore (m)
Re = Reynolds number (dimensionless)
However, Equation 10.2.3 is difficult to use because the friction factor appears on both sides of the
equation, and it is for this reason that manual calculations are likely to be carried out by using the Moody
chart.

On an SI style Moody chart, the friction factor scale might typically range from 0.002 to 0.02, whereas on
an Imperial style Moody chart, this scale might range from 0.008 to 0.08.

As a general rule, for turbulent flow with Reynolds numbers between 4000 and 100000, 'SI based' friction
factors will be of the order suggested by Equation 10.2.4, whilst 'Imperial based' friction factors will be of
the order suggested by Equation 10.2.5.

Equation 10.2.4 - 'SI based' friction factors

Equation 10.2.5 - 'Imperial based' friction factors


The friction factor used will determine whether the D'Arcy Equation 10.2.1 or 10.2.2 is used.
For 'SI based' friction factors, use Equation 10.2.1; for 'Imperial based' friction factors, use
Equation 10.2.2.

Example 10.2.1 - Water pipe


Determine the velocity, friction factor and the difference in pressure between two points
1 km apart in a 150 mm constant bore horizontal pipework system if the water flowrate is
45 m3/h at 15°C.
In essence, the friction factor depends on the Reynolds number (Re) of the flowing liquid and the relative
roughness (kS/d) of the inside of the pipe; the former calculated from Equation 10.2.6, and the latter from
Equation 10.2.7.

Reynolds number (Re)

Equation 10.2.6
Where:
Re = Reynolds number
ρ = Density of water = 1000 kg/m3
u = Velocity of water = 0.71 m/s
D = Pipe diameter = 0.15 m
 = Dynamic viscosity of water (at 15°C) = 1.138 x 10-3 kg/m s (from steam tables)
From Equation 10.2.6:

The pipe roughness or 'ks' value (often quoted as 'e' in some texts) is taken from standard tables, and for
'commercial steel pipe' would generally be taken as 0.000045 metres.

From this the relative roughness is determined (as this is what the Moody chart requires).

Equation 10.2.7
From Equation 10.2.7:

The friction factor can now be determined from the Moody chart and the friction head loss calculated from
the relevant D'Arcy Equation.

From the European Moody chart (Figure 10.2.4),


Where: ks/D = 0.0003 Re = 93585: Friction factor (f) = 0.005
Fig. 10.2.4 'SI based' Moody chart (abridged)
From the European D'Arcy equation (Equation 10.2.1):

From the USA / AUS Moody chart (Figure 10.2.5),


Where: ks/D = 0.0003 Re = 93 585 Friction factor (f) = 0.02
Fig. 10.2.5 'Imperial based' Moody chart (abridged)
From the USA / AUS D'Arcy equation (Equation 10.2.2):

The same friction head loss is obtained by using the different friction factors and relevant D'Arcy
equations.

In practice whether for water pipes or steam pipes, a balance is drawn between pipe size and pressure
loss.

Top
Steam

Oversized pipework means:


 Pipes, valves, fittings, etc. will be more expensive than necessary.
 Higher installation costs will be incurred, including support work, insulation, etc.
 For steam pipes a greater volume of condensate will be formed due to the greater heat
loss. This, in turn, means that either:
 More steam trapping is required, or
 Wet steam is delivered to the point of use.
In a particular example:

 The cost of installing 80 mm steam pipework was found to be 44% higher than the cost
of 50 mm pipework, which would have had adequate capacity.
 The heat lost by the insulated pipework was some 21% higher from the 80 mm pipeline
than it would have been from the 50 mm pipework. Any non-insulated parts of the 80 mm pipe
would lose 50% more heat than the 50 mm pipe, due to the extra heat transfer surface area.

Undersized pipework means:


 A lower pressure may only be available at the point of use. This may hinder equipment
performance due to only lower pressure steam being available.
 There is a risk of steam starvation.
 There is a greater risk of erosion, waterhammer and noise due to the inherent increase
in steam velocity.
As previously mentioned, the friction factor (f) can be difficult to determine, and the calculation itself is time
consuming especially for turbulent steam flow. As a result, there are numerous graphs, tables and slide
rules available for relating steam pipe sizes to flowrates and pressure drops.

One pressure drop sizing method, which has stood the test of time, is the 'pressure factor' method. A table
of pressure factor values is used in Equation 10.2.2 to determine the pressure drop for a particular
installation.

Equation 10.2.8
Where:
F = Pressure factor
P1 = Factor at inlet pressure
P2 = Factor at a distance of L metres
L = Equivalent length of pipe (m)

Example 10.2.2
Consider the system shown in Figure 10.2.6, and determine the pipe size required from the boiler
to the unit heater branch line. Unit heater steam load = 270 kg/h.
Fig. 10.2.6
System used to illustrate Example 10.2.2
Although the unit heater only requires 270 kg/h, the boiler has to supply more than this due to heat losses
from the pipe.

The allowance for pipe fittings


The length of travel from the boiler to the unit heater is known, but an allowance must be included for the
additional frictional resistance of the fittings. This is generally expressed in terms of 'equivalent pipe
length'. If the size of the pipe is known, the resistance of the fittings can be calculated. As the pipe size is
not yet known in this example, an addition to the equivalent length can be used based on experience.

 If the pipe is less than 50 metres long, add an allowance for fittings of 5%.
 If the pipe is over 100 metres long and is a fairly straight run with few fittings, an
allowance for fittings of 10% would be made.
 A similar pipe length, but with more fittings, would increase the allowance towards 20%.
In this instance, revised length = 150 m + 10% = 165 m

The allowance for the heat losses from the pipe


The unit heater requires 270 kg/h of steam; therefore the pipe must carry this quantity plus the quantity of
steam condensed by heat losses from the main. As the size of the main is yet to be determined, the true
calculations cannot be made, but, assuming that the main is insulated, it may be reasonable to add 3.5%
of the steam load per 100 m of the revised length as heat losses.

In this instance, the additional allowance =

Revised boiler load = 270 kg/h + 5.8% = 286 kg/h

From Table 10.2.2 (an extract from the complete pressure factor table, Table 10.2.5, which can be found in
the Appendix at the end of this Tutorial) 'F' can be determined by finding the pressure factors P 1 and P2,
and substituting them into Equation 10.2.8.

Table 10.2.2 Extract from pressure factor table (Table 10.2.5)


From the pressure factor table (see Table 10.2.2):

P1 (7.0 bar g) = 56.38

P2 (6.6 bar g) = 51.05

Substituting these pressure factors (P1 and P2) into Equation 10.2.8 will determine the value for 'F':

Equation 10.2.8.

Following down the left-hand column of the pipeline capacity and pressure drop factors table (Table 10.2.6
- Extract shown in Table 10.2.3); the nearest two readings around the requirement of 0.032 are 0.030 and
0.040. The next lower factor is always selected; in this case, 0.030.

Table 10.2.3 Extract from pipeline capacity and pressure factor table (Table 10.2.6)
Although values can be interpolated, the table does not conform exactly to a straight-line graph, so
interpolation cannot be absolutely correct. Also, it is bad practice to size any pipe up to the limit of its
capacity, and it is important to have some leeway to allow for the inevitable future changes in design.

From factor 0.030, by following the row of figures to the right it will be seen that:

 A 40 mm pipe will carry 229.9 kg/h.


 A 50 mm pipe will carry 501.1 kg/h.
Since the application requires 286 kg/h, the 50 mm pipe would be selected.

Having sized the pipe using the pressure drop method, the velocity can be checked if required.

Where:

Viewed in isolation, this velocity may seem low in comparison with maximum permitted velocities.
However, this steam main has been sized to limit pressure drop, and the next smaller pipe size would have
given a velocity of over 47 m/s, and a final pressure less than the requirement of 6.6 bar g, which is
unacceptable.

As can be seen, this procedure is fairly complex and can be simplified by using the nomogram shown in
Figure 10.2.9 (in the Appendix of this Tutorial). The method of use is explained in Example 10.2.3.

Example 10.2.3
Using the data from Example 10.2.2, determine the pressure drop using the nomogram shown in
Figure 10.2.7.

Inlet pressure = 7 bar g

Steam flowrate = 286 kg/h

Minimum allowable P2 = 6.6 bar g

Method

 Select the point on the saturated steam line at 7 bar g, and mark Point A.
 From point A, draw a horizontal line to the steam flowrate of 286 kg/h, and mark Point B.
 From point B, draw a vertical line towards the top of the nomogram (Point C).
 Draw a horizontal line from 0.24 bar/100 m on the pressure loss scale (Line DE).
 The point at which lines DE and BC cross will indicate the pipe size required. In this
case, a 40 mm pipe is too small, and a 50 mm pipe would be used.
Fig. 10.2.7 Steam pipeline sizing chart - Pressure drop

Sizing pipes on velocity


From the knowledge gained at the beginning of this Tutorial, and particularly the notes regarding the
D'Arcy equation (Equation 10.2.1), it is acknowledged that velocity is an important factor in sizing pipes. It
follows then, that if a reasonable velocity could be used for a particular fluid flowing through pipes, then
velocity could be used as a practical sizing factor. As a general rule, a velocity of 25 to 40 m/s is used
when saturated steam is the medium.

40 m/s should be considered an extreme limit, as above this, noise and erosion will take place particularly
if the steam is wet.

Even these velocities can be high in terms of their effect on pressure drop. In longer supply lines, it is often
necessary to restrict velocities to 15 m/s to avoid high pressure drops. It is recommended that pipelines
over 50 m long are always checked for pressure drop, no matter what the velocity.

By using Table 10.2.4 as a guide, it is possible to select pipe sizes from known data; steam pressure,
velocity and flowrate.
Table 10.2.4 Saturated steam pipeline capacities in kg/h for different velocities (Schedule 40 pipe)
Alternatively the pipe size can be calculated arithmetically. The following information is required, and the
procedure used for the calculation is outlined below.

Information required to calculate the required pipe size:

u = Flow velocity (m/s)

vg = Specific volume (m3/kg)

s = Mass flowrate (kg/s)

= Volumetric flowrate (m3/s) = ms x vg

From this information, the cross sectional area (A) of the pipe can be calculated:
Rearranging the formula to give the diameter of the pipe (D) in metres:

Example 10.2.4
A process requires 5 000 kg/h of dry saturated steam at 7 bar g. For the flow velocity not to exceed
25 m/s, determine the pipe size.

Where

Therefore, using:

Since the steam velocity must not exceed 25 m/s, the pipe size must be at least 130 mm; the
nearest commercially available size, 150 mm, would be selected.

Again, a nomogram has been created to simplify this process, see Figure 10.2.6.

Example 10.2.5
Using the information from Example 10.2.4, use Figure 10.2.6 to determine the minimum
acceptable pipe size

Inlet pressure = 7 bar g


Steam flowrate = 5000 kg/h
Maximum velocity = 25 m/s
Method:

 Draw a horizontal line from the saturation temperature line at 7 bar g (Point A) on the
pressure scale to the steam mass flowrate of 5 000 kg/h (Point B).
 From point B, draw a vertical line to the steam velocity of 25 m/s (Point C). From point C,
draw a horizontal line across the pipe diameter scale (Point D).
 A pipe with a bore of 130 mm is required; the nearest commercially available size, 150
mm, would be selected.

Fig. 10.2.8 Steam pipeline sizing chart - Velocity

Sizing pipes for superheated steam duty


Superheated steam can be considered as a dry gas and therefore carries no moisture. Consequently there
is no chance of pipe erosion due to suspended water droplets, and steam velocities can be as high as 50
to 70 m/s if the pressure drop permits this. The nomograms in Figures 10.2.5 and 10.2.6 can also be used
for superheated steam applications.

Example 10.2.6
Utilising the waste heat from a process, a boiler/superheater generates 30 t/h of superheated steam
at 50 bar g and 450°C for export to a neighbouring power station. If the velocity is not to exceed 50
m/s, determine:

1. The pipe size based on velocity (use Figure 10.2.8).


2. The pressure drop if the pipe length, including allowances, is 200 m (use Figure
10.2.9).

Part 1

 Using Figure 10.2.8, draw a vertical line from 450°C on the temperature axis until it
intersects the 50 bar line (Point A).
 From point A, project a horizontal line to the left until it intersects the steam 'mass
flowrate' scale of 30 000 kg/h (30 t/h) (Point B).
 From point B, project a line vertically upwards until it intersects 50 m/s on the 'steam
velocity' scale (Point C).
 From Point C, project a horizontal line to the right until it intersects the 'inside pipe
diameter' scale.
The 'inside pipe diameter' scale recommends a pipe with an inside diameter of about 120 mm. From Table
10.2.1 and assuming that the pipe will be Schedule 80 pipe, the nearest size would be 150 mm, which has
a bore of 146.4 mm.

Part 2

 Using Figure 10.2.7, draw a vertical line from 450°C on the temperature axis until it
intersects the 50 bar line (Point A).
 From point A, project a horizontal line to the right until it intersects the 'steam mass
flowrate' scale of 30 000 kg/h (30 t/h) (Point B).
 From point B, project a line vertically upwards until it intersects the 'inside pipe diameter'
scale of (approximately) 146 mm (Point C).
 From Point C, project a horizontal line to the left until it intersects the 'pressure loss
bar/100 m' scale (Point D).
The 'pressure loss bar/100 m' scale reads about 0.9 bar/100 m. The pipe length in the example is 200 m,
so the pressure drop is:

This pressure drop must be acceptable at the process plant.

Using formulae to establish steam flowrate on pressure drop


Empirical formulae exist for those who prefer to use them. Equations 10.2.9 and 10.2.10 are shown below.
These have been tried and tested over many years, and which appear to give results close to the pressure
factor method. The advantage of using these formulae is that they can be programmed into a scientific
calculator, or a spreadsheet, and consequently used without the need to look up tables and charts.
Equation 10.2.10 requires the specific volume of steam to be known, which means it is necessary to look
up this value from a steam table. Also, Equation 10.2.10 should be restricted to a maximum pipe length of
200 metres.

Pressure drop formula 1

Equation 10.2.9
Where:
P1 = Upsteam pressure (bar a)
P2 = Downstream pressure (bar a)
L = Length of pipe (m)
s = Mass flowrate (kg/h)

D = Pipe diameter (mm)


Pressure drop formula 2 (Maximum pipe length: 200 metres)
Equation 10.2.10
Where:
P = Pressure drop (bar)
L = Length of pipe
vg = Specific volume of steam (m3/kg)
= Mass flowrate(kg/h)
D = Pipe diameter (mm)
Top

Summary
 The selection of piping material and the wall thickness required for a particular
installation is stipulated in standards such as EN 45510 and ASME 31.1.
 Selecting the appropriate pipe size (nominal bore) for a particular application is based on
accurately identifying pressure and flowrate. The pipe size may be selected on the basis of:
 Velocity (usually pipes less than 50 m in length).
 Pressure drop (as a general rule, the pressure drop should not normally exceed 0.1
bar/50 m.
Top

Appendix
Table 10.2.5 Pressure drop factor (F) table
Table 10.2.6 Pipeline capacity and pressure factor table
Fig 10.2.9 Steam pipeline sizing chart - Pressure drop
Fig 10.2.10 Steam pipeline sizing chart - Velocity
Top
Steam Mains and Drainage
Issues surrounding the structure, layout and operation of a steam distribution system, including
condensate drain points and branch lines, the avoidance of waterhammer and the use of
separators and strainers for steam conditioning.

Use the quick links below to take you to the main sections of this tutorial:

 Piping layout
 Waterhammer and its effects
 Branch lines
 Rising ground and drainage
 Steam separators
 Strainers
 How to drain steam mains
 Steam leaks
 Summary
 The printable version of this page has now been replaced by
The Steam and Condensate Loop Book
 Try answering the Questions for this tutorial
 View the complete collection of Steam Engineering Tutorials
 Contact Us

Throughout the length of a hot steam main, an amount of heat will be transferred to the environment, and
this will depend on the parameters identified in Block 2 - 'Steam Engineering and Heat Transfer', and
brought together in Equation 2.5.1.

Equation 2.5.1
Where:
= Heat transferred per unit time (W)
k = Thermal conductivity of the material (W/m K or W/m °C)
A = Heat transfer area (m2)
T = Temperature difference across the material (K or °C)
 = Material thickness (m)

With steam systems, this loss of energy represents inefficiency, and thus pipes are insulated to limit these
losses. Whatever the quality or thickness of insulation, there will always be a level of heat loss, and this
will cause steam to condense along the length of the main.

The effect of insulation is discussed in Tutorial 10.5. This Tutorial will concentrate on disposal of the
inevitable condensate, which, unless removed, will accumulate and lead to problems such as corrosion,
erosion, and waterhammer.

In addition, the steam will become wet as it picks up water droplets, which reduces its heat transfer
potential. If water is allowed to accumulate, the overall effective cross sectional area of the pipe is reduced,
and steam velocity can increase above the recommended limits.

Top

Piping layout
The subject of drainage from steam lines is covered in the European Standard EN 45510, Section 4.12.

EN 45510 states that, whenever possible, the main should be installed with a fall of not less than 1:100 (1
m fall for every 100 m run), in the direction of the steam flow. This slope will ensure that gravity, as well as
the flow of steam, will assist in moving the condensate towards drain points where the condensate may be
safely and effectively removed (See Figure 10.3.1).

Fig.
10.3.1 Typical steam main installation

Drain points
The drain point must ensure that the condensate can reach the steam trap. Careful consideration must
therefore be given to the design and location of drain points.

Consideration must also be given to condensate remaining in a steam main at shutdown, when steam flow
ceases. Gravity will ensure that the water (condensate) will run along sloping pipework and collect at low
points in the system. Steam traps should therefore be fitted to these low points.

The amount of condensate formed in a large steam main under start-up conditions is sufficient to require
the provision of drain points at intervals of 30 m to 50 m, as well as natural low points such as at the
bottom of rising pipework.

In normal operation, steam may flow along the main at speeds of up to 145 km/h, dragging condensate
along with it. Figure 10.3.2 shows a 15 mm drain pipe connected directly to the bottom of a main.

Fig. 10.3.2 Trap


pocket too small
Although the 15 mm pipe has sufficient capacity, it is unlikely to capture much of the condensate moving
along the main at high speed. This arrangement will be ineffective.

A more reliable solution for the removal of condensate is shown in Figure 10.3.3. The trap line should be at
least 25 to 30 mm from the bottom of the pocket for steam mains up to 100 mm, and at least 50 mm for
larger mains. This allows a space below for any dirt and scale to settle.

Fig. 10.3.3 Trap


pocket properly sized
The bottom of the pocket may be fitted with a removable flange or blowdown valve for cleaning purposes.

Recommended drain pocket dimensions are shown in Table 10.3.1 and in Figure 10.3.4.
Table 10.3.1 Recomended drain pocket dimensions
Top

Waterhammer and its effects


Waterhammer is the noise caused by slugs of condensate colliding at high velocity into pipework fittings,
plant, and equipment. This has a number of implications:

 Because the condensate velocity is higher than normal, the dissipation of kinetic energy
is higher than would normally be expected.
 Water is dense and incompressible, so the 'cushioning' effect experienced when gases
encounter obstructions is absent.
 The energy in the water is dissipated against the obstructions in the piping system such
as valves and fittings.

Fig.
10.3.5 Formation of a 'solid' slug of water
Indications of waterhammer include a banging noise, and perhaps movement of the pipe.

In severe cases, waterhammer may fracture pipeline equipment with almost explosive effect, with
consequent loss of live steam at the fracture, leading to an extremely hazardous situation.

Good engineering design, installation and maintenance will avoid waterhammer; this is far better practice
than attempting to contain it by choice of materials and pressure ratings of equipment.

Commonly, sources of waterhammer occur at the low points in the pipework (See Figure 10.3.6). Such
areas are due to:

 Sagging in the line, perhaps due to failure of supports.


 Incorrect use of concentric reducers (see Figure 10.3.7) - Always use eccentric reducers
with the flat at the bottom.
 Incorrect strainer installation - They should be fitted with the basket on the side.
 Inadequate drainage of steam lines.
 Incorrect operation - Opening valves too quickly at start-up when pipes are cold.

Fig.
10.3.6 Potential sources of waterhammer

Fig. 10.3.7 Eccentric and concentric pipe


reducers
To summarise, the possibility of waterhammer is minimised by:

 Installing steam lines with a gradual fall in the direction of flow, and with drain points
installed at regular intervals and at low points.
 Installing check valves after all steam traps which would otherwise allow condensate to
run back into the steam line or plant during shutdown.
 Opening isolation valves slowly to allow any condensate which may be lying in the
system to flow gently through the drain traps, before it is picked up by high velocity steam. This is
especially important at start-up.
Top

Branch lines
Fig. 10.3.8
Branch line
Branch lines are normally much shorter than steam mains. As a general rule, therefore, provided the
branch line is not more than 10 metres in length, and the pressure in the main is adequate, it is possible to
size the pipe on a velocity of 25 to 40 m/s, and not to worry about the pressure drop.

Table 10.2.4 'Saturated steam pipeline capacities for different velocities' in Tutorial 10.2 will prove useful in
this exercise.

Branch line connections


Branch line connections taken from the top of the main carry the driest steam (Figure 10.3.8). If
connections are taken from the side, or even worse from the bottom (as in Figure 10.3.9 (a)), they can
accept the condensate and debris from the steam main. The result is very wet and dirty steam reaching
the equipment, which will affect performance in both the short and long term.

The valve in Figure 10.3.9 (b) should be positioned as near to the off-take as possible to minimise
condensate lying in the branch line, if the plant is likely to be shutdown for any extended periods.
Fig. 10.3.9 Steam off-take

Drop leg
Low points will also occur in branch lines. The most common is a drop leg close to an isolating valve or a
control valve (Figure 10.3.10). Condensate can accumulate on the upstream side of the closed valve, and
then be propelled forward with the steam when the valve opens again - consequently a drain point with a
steam trap set is good practice just prior to the strainer and control valve.

Fig. 10.3.10 Diagram of a drop leg supplying a unit heater


Top

Rising ground and drainage


There are many occasions when a steam main must run across rising ground, or applications where the
contours of the site make it impractical to lay the pipe with the 1:100 fall proposed earlier. In these
situations, the condensate must be encouraged to run downhill and against the steam flow. Good practice
is to size the pipe on a low steam velocity of not more than 15 m/s, to run the line at a slope of no less than
1:40, and install the drain points at not more than 15 metre intervals (see Figure 10.3.11).
The objective is to prevent the condensate film on the bottom of the pipe increasing in thickness to the
point where droplets can be picked up by the steam flow.

Fig. 10.3.11
Reverse gradient on steam main
Top

Steam separators
Modern packaged steam boilers have a large evaporating capacity for their size and have limited capacity
to cope with rapidly changing loads. In addition, as discussed in Block 3 'The Boiler House', other
circumstances, such as . . .

 Incorrect chemical feedwater treatment and/or TDS control


 Transient peak loads in other parts of the plant
. . . can cause priming and carryover of boiler water into the steam mains.

Separators, as shown by the cut section in Figure 10.3.12, may be installed to remove this water.

Fig. 10.3.12 Cut section


through a separator
As a general rule, providing the velocities in the pipework are within reasonable limits, separators will be
line sized. (Separators are discussed in detail in Tutorial 12.5)
A separator will remove both droplets of water from pipe walls and suspended mist entrained in the steam
itself. The presence and effect of waterhammer can be eradicated by fitting a separator in a steam main,
and can often be less expensive than increasing the pipe size and fabricating drain pockets.

A separator is recommended before control valves and flowmeters. It is also wise to fit a separator where a
steam main enters a building from outside. This will ensure that any condensate produced in the external
distribution system is removed and the building always receives dry steam. This is equally important where
steam usage in the building is monitored and charged for.

Top

Strainers
When new pipework is installed, it is not uncommon for fragments of casting sand, packing, jointing, swarf,
welding rods and even nuts and bolts to be accidentally deposited inside the pipe. In the case of older
pipework, there will be rust, and in hard water districts, a carbonate deposit. Occasionally, pieces will break
loose and pass along the pipework with the steam to rest inside a piece of steam using equipment. This
may, for example, prevent a valve from opening/closing correctly. Steam using equipment may also suffer
permanent damage through wiredrawing - the cutting action of high velocity steam and water passing
through a partly open valve. Once wiredrawing has occurred, the valve will never give a tight shut-off, even
if the dirt is removed.

It is therefore wise to fit a line-size strainer in front of every steam trap, flowmeter, reducing valve and
regulating valve. The illustration shown in Figure 10.3.13 shows a cut section through a typical strainer.

Fig. 10.3.13 Cut section


through a Y-type strainer.
Steam flows from the inlet 'A' through the perforated screen 'B' to the outlet 'C'. While steam and water
will pass readily through the screen, dirt cannot. The cap 'D', can be removed, allowing the screen to be
withdrawn and cleaned at regular intervals. A blowdown valve can also be fitted to cap 'D' to facilitate
regular cleaning.

Strainers can however, be a source of wet steam as previously mentioned. To avoid this situation, strainers
should always be installed in steam lines with their baskets to the side.

Strainers and screen details are discussed in Tutorial 12.4.

Top

How to drain steam mains


Steam traps are the most effective and efficient method of draining condensate from a steam distribution
system.

The steam traps selected must suit the system in terms of:

 Pressure rating
 Capacity
 Suitability

Pressure rating
Pressure rating is easily dealt with; the maximum possible working pressure at the steam trap will either be
known or should be established.

Capacity
Capacity, that is, the quantity of condensate to be discharged, which needs to be divided into two
categories; warm-up load and running load.

Warm-up load - In the first instance, the pipework needs to be brought up to operating temperature. This
can be determined by calculation, knowing the mass and specific heat of the pipework and fittings.
Alternatively, Table 10.3.2 may be used.

 The table shows the amount of condensate generated when bringing 50 m of steam
main up to working temperature; 50 m being the maximum recommended distance between
trapping points.
 The values shown are in kilograms. To determine the average condensing rate, the time
taken for the process must be considered. For example, if the warm-up process required 50 kg of
steam, and was to take 20 minutes, then the average condensing rate would be:

 When using these capacities to size a steam trap, it is worth remembering that the initial
pressure in the main will be little more than atmospheric when the warm-up process begins.
However, the condensate loads will still generally be well within the capacity of a DN15 'low
capacity' steam trap. Only in rare applications at very high pressures (above 70 bar g), combined
with large pipe sizes, will greater trap capacity be needed.
Running load - Once the steam main is up to operating temperature, the rate of condensation is mainly a
function of the pipe size and the quality and thickness of the insulation.

Alternatively, for quick approximations of running load, Table 10.3.3 can be used which shows typical
amounts of steam condensed each hour per 50 m of insulated steam main at various pressures. For
accurate means of calculating running losses from steam mains, refer to Tutorial 2.12 'Steam consumption
of pipes and air heaters'.
Table 10.3.2 Amount of steam condensed to warm-up 50 m of schedule 40 pipe (kg)
Note: Figures are based on an ambient temperature of 20°C, and an insulation efficiency of 80%

Table 10.3.3 Condensing rate of steam in 50 m of schedule 40 pipe - at working temperature (kg/h)
Note: Figures are based on an ambient temperature of 20°C, and an insulation efficiency of 80%

Suitability
A mains drain trap should consider the following constraints:

 Discharge temperature - The steam trap should discharge at, or very close to
saturation temperature, unless cooling legs are used between the drain point and the trap. This
means that the choice is a mechanical type trap (such as a float, inverted bucket type, or
thermodynamic traps).
 Frost damage - Where the steam main is located outside a building and there is a
possibility of sub-zero ambient temperature, the thermodynamic steam trap is ideal, as it is not
damaged by frost. Even if the installation causes water to be left in the trap at shutdown and
freezing occurs, the thermodynamic trap may be thawed out without suffering damage when
brought back into use.
 Waterhammer - In the past, on poorly laid out installations where waterhammer was a
common occurrence, float traps were not always ideal due to their susceptibility to float damage.
Contemporary design and manufacturing techniques now produce extremely robust units for
mains drainage purposes. Float traps are certainly the first choice for proprietary separators as
high capacities are readily achieved, and they are able to respond quickly to rapid load increases.
Steam traps used to drain condensate from steam mains, are shown in Figure 10.3.14. The thermostatic
trap is included because it is ideal where there is no choice but to discharge condensate into a flooded
return pipe.

The subject of steam trapping is dealt with in detail in the Block 11, 'Steam Trapping'.

Fig. 10.3.14 Steam traps suitable for steam mains drainage


Top

Steam leaks
Steam leaking from pipework is often ignored. Leaks can be costly in both the economic and
environmental sense and therefore need prompt attention to ensure the steam system is working at its
optimum efficiency with a minimum impact on the environment.

Figure 10.3.15 illustrates the steam loss for various sizes of hole at various pressures. This loss can be
readily translated into a fuel saving based on the annual hours of operation.
Fig. 10.3.15 Steam
leakage rate through holes
Top

Summary
Proper pipe alignment and drainage means observing a few simple rules:

 Steam lines should be arranged to fall in the direction of flow, at not less than 100 mm
per 10 metres of pipe (1:100). Steam lines rising in the direction of flow should slope at not less
than 250 mm per 10 metres of pipe (1:40).
 Steam lines should be drained at regular intervals of 30-50 m and at any low points in
the system.
 Where drainage has to be provided in straight lengths of pipe, then a large bore pocket
should be used to collect condensate.
 If strainers are to be fitted, then they should be fitted on their sides.
 Branch connections should always be taken from the top of the main from where the
driest steam is taken.
 Separators should be considered before any piece of steam using equipment ensuring
that dry steam is used.
 Traps selected should be robust enough to avoid waterhammer damage and frost
damage.
Top
Pipe Expansion and Support
Any steam system must be fully supported, able to expand during operation and sufficiently
flexible to allow movement as a result. This tutorial includes advice on different methods and full
calculations.

Use the quick links below to take you to the main sections of this tutorial:

 Allowance for expansion


 Pipework flexibility
 Expansion fittings
 Pipe support spacing
 The printable version of this page has now been replaced by
The Steam and Condensate Loop Book
 Try answering the Questions for this tutorial
 View the complete collection of Steam Engineering Tutorials
 Contact Us

Top

Allowance for expansion

Allowance for expansion


All pipes will be installed at ambient temperature. Pipes carrying hot fluids such as water or steam operate
at higher temperatures.

It follows that they expand, especially in length, with an increase from ambient to working temperatures.
This will create stress upon certain areas within the distribution system, such as pipe joints, which, in the
extreme, could fracture. The amount of the expansion is readily calculated using Equation 10.4.1, or read
from an appropriate chart such as Figure 10.4.1.

Equation 10.4.1
Where:
L = Length of pipe between anchors (m)
T = Temperature difference between ambient temperature and operating temperatures (°C)
 = Expansion coefficient (mm/m °C) x 10-3

Table 10.4.1 Expansion coefficients (a) (mm/m °C x 10-3)


Example 10.4.1
A 30 m length of carbon steel pipe is to be used to transport steam at 4 bar g (152°C). If the pipe is
installed at 10°C, determine the expansion using Equation 10.4.1.
Alternatively, the chart in Figure 10.4.1 can be used for finding the approximate expansion of a variety of
steel pipe lengths - see Example 10.4.2 for explanation of use.

Example 10.4.2
Using Figure 10.4.1. Find the approximate expansion from 15°C, of 100 metres of carbon steel pipework
used to distribute steam at 265°C.

Temperature difference is 265 - 15°C = 250°C.

Where the diagonal temperature difference line of 250°C cuts the horizontal pipe length line at 100 m, drop
a vertical line down. For this example an approximate expansion of 330 mm is indicated.

Fig. 10.4.1 A
chart showing the expansion in various steel pipe lengths at various temperature differences

Table 10.4.2 Temperature of saturated steam


Top

Pipework flexibility
The pipework system must be sufficiently flexible to accommodate the movements of the components as
they expand. In many cases the flexibility of the pipework system, due to the length of the pipe and
number of bends and supports, means that no undue stresses are imposed. In other installations,
however, it will be necessary to incorporate some means of achieving this required flexibility.

An example on a typical steam system is the discharge of condensate from a steam mains drain trap into
the condensate return line that runs along the steam line (Figure 10.4.2). Here, the difference between the
expansions of the two pipework systems must be taken into account. The steam main will be operating at
a higher temperature than that of the condensate main, and the two connection points will move relative to
each other during system warm-up.

Fig. 10.4.2 Flexibility in


connection to condensate return line
The amount of movement to be taken up by the piping and any device incorporated in it can be reduced by
'cold draw'. The total amount of expansion is first calculated for each section between fixed anchor points.
The pipes are left short by half of this amount, and stretched cold by pulling up bolts at a flanged joint, so
that at ambient temperature, the system is stressed in one direction. When warmed through half of the
total temperature rise, the piping is unstressed. At working temperature and having fully expanded, the
piping is stressed in the opposite direction. The effect is that instead of being stressed from 0 F to +1 F
units of force, the piping is stressed from -½ F to + ½ F units of force.

In practical terms, the pipework is assembled cold with a spacer piece, of length equal to half the
expansion, between two flanges. When the pipework is fully installed and anchored at both ends, the
spacer is removed and the joint pulled up tight (see Figure 10.4.3).

Fig. 10.4.3 Use


of spacer for expansion when pipework is installed
The remaining part of the expansion, if not accepted by the natural flexibility of the pipework will call for the
use of an expansion fitting.

In practice, pipework expansion and support can be classified into three areas as shown in Figure 10.4.4.

Fi
g. 10.4.4 Diagram of pipeline with fixed point, variable anchor point and expansion fitting
The fixed or 'anchor' points 'A' provide a datum position from which expansion takes place.

The sliding support points 'B' allow free movement for expansion of the pipework, while keeping the
pipeline in alignment.

The expansion device at point 'C' is to accommodate the expansion and contraction of the pipe.

Fig. 10.4.5 Chair and roller

Fig. 10.4.6 Chair roller and saddle


Roller supports (Figure 10.4.5 and 10.4.6) are ideal methods for supporting pipes, at the same time
allowing them to move in two directions. For steel pipework, the rollers should be manufactured from
ferrous material. For copper pipework, they should be manufactured from non-ferrous material. It is good
practice for pipework supported on rollers to be fitted with a pipe saddle bolted to a support bracket at not
more than distances of 6 metres to keep the pipework in alignment during any expansion and contraction.

Where two pipes are to be supported one below the other, it is poor practice to carry the bottom pipe from
the top pipe using a pipe clip. This will cause extra stress to be added to the top pipe whose thickness has
been sized to take only the stress of its working pressure.

All pipe supports should be specifically designed to suit the outside diameter of the pipe concerned.

Top

Expansion fittings
The expansion fitting ('C' Figure 10.4.4) is one method of accommodating expansion. These fittings are
placed within a line, and are designed to accommodate the expansion, without the total length of the line
changing. They are commonly called expansion bellows, due to the bellows construction of the expansion
sleeve.

Other expansion fittings can be made from the pipework itself. This can be a cheaper way to solve the
problem, but more space is needed to accommodate the pipe.

Full loop
This is simply one complete turn of the pipe and, on steam pipework, should preferably be fitted in a
horizontal rather than a vertical position to prevent condensate accumulating on the upstream side.

The downstream side passes below the upstream side and great care must be taken that it is not fitted the
wrong way round, as condensate can accumulate in the bottom. When full loops are to be fitted in a
confined space, care must be taken to specify that wrong-handed loops are not supplied.
The full loop does not produce a force in opposition to the expanding pipework as in some other types, but
with steam pressure inside the loop, there is a slight tendency to unwind, which puts an additional stress
on the flanges.

Fig. 10.4.7 Full loop


This design is used rarely today due to the space taken up by the pipework, and proprietary expansion
bellows are now more readily available. However large steam users such as power stations or
establishments with large outside distribution systems still tend to use full loop type expansion devices, as
space is usually available and the cost is relatively low.

Horseshoe or lyre loop


When space is available this type is sometimes used. It is best fitted horizontally so that the loop and the
main are on the same plane. Pressure does not tend to blow the ends of the loop apart, but there is a very
slight straightening out effect. This is due to the design but causes no misalignment of the flanges.

If any of these arrangements are fitted with the loop vertically above the pipe then a drain point must be
provided on the upstream side as depicted in Figure 10.4.8.

Fig. 10.4.8 Horseshoe or lyre loop

Expansion loops

Fig. 10.4.9
Expansion loop
The expansion loop can be fabricated from lengths of straight pipes and elbows welded at the joints
(Figure 10.4.9). An indication of the expansion of pipe that can be accommodated by these assemblies is
shown in Figure 10.4.10.

It can be seen from Figure 10.4.9 that the depth of the loop should be twice the width, and the width is
determined from Figure 10.4.10, knowing the total amount of expansion expected from the pipes either
side of the loop.

Fi
g. 10.4.10 Expansion loop capacity for carbon steel pipes

Sliding joint
These are sometimes used because they take up little room, but it is essential that the pipeline is rigidly
anchored and guided in strict accordance with the manufacturers' instructions; otherwise steam pressure
acting on the cross sectional area of the sleeve part of the joint tends to blow the joint apart in opposition
to the forces produced by the expanding pipework (see Figure 10.4.11). Misalignment will cause the sliding
sleeve to bend, while regular maintenance of the gland packing may also be needed.
Fig. 10.4.11
Sliding joint

Expansion bellows
An expansion bellows, Figures 10.4.12, has the advantage that it requires no packing (as does the sliding
joint type). But it does have the same disadvantages as the sliding joint in that pressure inside tends to
extend the fitting, consequently, anchors and guides must be able to withstand this force.

Fig. 10.4.12 Simple expansion bellows


Bellows may incorporate limit rods, which limit over-compression and over-extension of the element.
These may have little function under normal operating conditions, as most simple bellows assemblies are
able to withstand small lateral and angular movement. However, in the event of anchor failure, they
behave as tie rods and contain the pressure thrust forces, preventing damage to the unit whilst reducing
the possibility of further damage to piping, equipment and personnel (Figure 10.4.13 (b)).

Where larger forces are expected, some form of additional mechanical reinforcement should be built into
the device, such as hinged stay bars (Figure 10.4.13 (c)).

There is invariably more than one way to accommodate the relative movement between two laterally
displaced pipes depending upon the relative positions of bellows anchors and guides. In terms of
preference, axial displacement is better than angular, which in turn, is better than lateral. Angular and
lateral movement should be avoided wherever possible.

Figure 10.4.13 (a), (b), and (c) give a rough indication of the effects of these movements, but, under all
circumstances, it is highly recommended that expert advice is sought from the bellows' manufacturer
regarding any installation of expansion bellows.
Fig. 10.4.13 (a) Axial movement of bellows

Fig. 10.4.13 (b) Lateral and angular movement of bellows

Fig. 10.4.13
(c) Angular and axial movement of bellows
Top

Pipe support spacing


The frequency of pipe supports will vary according to the bore of the pipe; the actual pipe material (i.e.
steel or copper); and whether the pipe is horizontal or vertical.

Some practical points worthy of consideration are as follows:

 Pipe supports should be provided at intervals not greater than shown in Table 10.4.3,
and run along those parts of buildings and structures where appropriate supports may be
mounted.
 Where two or more pipes are supported on a common bracket, the spacing between the
supports should be that for the smallest pipe.
 When an appreciable movement will occur, i.e. where straight pipes are greater than 15
metres in length, the supports should be of the roller type as outlined previously.
 Vertical pipes should be adequately supported at the base, to withstand the total weight
of the vertical pipe and the fluid within it. Branches from vertical pipes must not be used as a
means of support for the pipe, because this will place undue strain upon the tee joint.
 All pipe supports should be specifically designed to suit the outside diameter of the pipe
concerned. The use of oversized pipe brackets is not good practice.
Table 10.4.3 can be used as a guide when calculating the distance between pipe supports for steel and
copper pipework.

Table 10.4.3 Recommended support for pipework


The subject of pipe supports is covered comprehensively in the European standard EN 13480,
Part3.

Top
Air Venting, Heat Losses and a Summary of Various Pipe
Related Standards
The venting of air and other incondensable gases from steam systems, and the provision of
adequate insulation, are vital to ensure steam plant efficiency, safety and performance.

Use the quick links below to take you to the main sections of this tutorial:

 Air venting
 Reduction of heat losses
 Calculation of heat transfer
 Relevant UK and International Standards
 Summary
 The printable version of this page has now been replaced by
The Steam and Condensate Loop Book
 Try answering the Questions for this tutorial
 View the complete collection of Steam Engineering Tutorials
 Contact Us

Top

Air venting
When steam is first admitted to a pipe after a period of shutdown, the pipe is full of air. Further amounts of
air and other non-condensable gases will enter with the steam, although the proportions of these gases
are normally very small compared with the steam. When the steam condenses, these gases will
accumulate in pipes and heat exchangers. Precautions should be taken to discharge them. The
consequence of not removing air is a lengthy warming up period, and a reduction in plant efficiency and
process performance.

Air in a steam system will also affect the system temperature. Air will exert its own pressure within the
system, and will be added to the pressure of the steam to give a total pressure. Therefore, the actual
steam pressure and temperature of the steam/air mixture will be lower than that suggested by a pressure
gauge.

Of more importance is the effect air has upon heat transfer. A layer of air only 1 mm thick can offer the
same resistance to heat as a layer of water 25 µm thick, a layer of iron 2 mm thick or a layer of copper 15
mm thick. It is very important therefore to remove air from any steam system.

Automatic air vents for steam systems (which operate on the same principle as thermostatic steam traps)
should be fitted above the condensate level so that only air or steam/air mixtures can reach them. The
best location for them is at the end of the steam mains as shown in Figure 10.5.1.
Fig. 10.5.1 Draining and
venting at the end of a steam main
The discharge from an air vent must be piped to a safe place. In practice, a condensate line falling towards
a vented receiver can accept the discharge from an air vent.

In addition to air venting at the end of a main, air vents should also be fitted:

 In parallel with an inverted bucket trap or, in some instances, a thermodynamic trap.
These traps are sometimes slow to vent air on start-up.
 In awkward steam spaces (such as at the opposite side to where steam enters a
jacketed pan).
 Where there is a large steam space (such as an autoclave), and a steam/air mixture
could affect the process quality.
Top

Reduction of heat losses


Even when a steam main has warmed up, steam will continue condensing as heat is lost by radiation. The
condensing rate will depend upon the steam temperature, the ambient temperature, and the efficiency of
the pipe insulation.

For a steam distribution system to be efficient, appropriate steps should be taken to ensure that heat
losses are reduced to the economic minimum. The most economical thickness of insulation will depend
upon several factors:

 Installation cost.
 The heat carried by the steam.
 Size of the pipework.
 Pipework temperature.
When insulating external pipework, dampness and wind speed must be taken into account.

The effectiveness of most insulation materials depends on minute air cells which are held in a matrix of
inert material such as mineral wool, fibreglass or calcium silicate. Typical installations use aluminium clad
fibreglass, aluminium clad mineral wool and calcium silicate. It is important that insulating material is not
crushed or allowed to waterlog. Adequate mechanical protection and waterproofing are essential,
especially in outdoor locations.

The heat loss from a steam pipe to water, or to wet insulation, can be as much as 50 times greater than
from the same pipe to air. Particular care should be taken to protect steam lines, running through
waterlogged ground, or in ducts, which may be subjected to flooding. The same applies to protecting the
lagging from damage by ladders etc., to avoid the ingress of rainwater.

It is important to insulate all hot parts of the system with the exception of safety valves. This includes all
flanged joints on the mains, and also the valves and other fittings. It was, at one time, common to cut back
the insulation at each side of a flanged joint, to leave access to the bolts for maintenance purposes. This is
equivalent to leaving about 0.5 m of bare pipe.

Fortunately, prefabricated insulating covers for flanged joints and valves are now more widely available.
These are usually provided with fasteners so that they can readily be detached to provide access for
maintenance purposes.

Top

Calculation of heat transfer


The calculation of heat losses from pipes can be very complex and time consuming, and assume that
obscure data concerning pipe wall thickness, heat transfer coefficients and various derived constants are
easily available, which, usually, they are not.

The derivations of these formulae are outside the scope of this Tutorial, but further information can be
readily found in any good thermodynamics textbook. To add to this, an abundance of contemporary
computer software is available for the discerning engineer.

This being so, pipe heat losses can easily be found by reference to Table 10.5.1 and a simple equation
(Equation 2.12.2).

The table assumes ambient conditions of between 10 - 21°C, and considers heat losses from bare
horizontal pipes of different sizes with steam contained at various pressures.

Table 10.5.1 Heat emission from pipes


Note: Heat emission from bare horizontal pipes with ambient temperatures between 10°C and 20°C and
still air conditions

Other factors can be included in the equation, for instance, if a pipe is lagged with insulation providing a
reduction in heat losses to 10% of the uninsulated pipe, then it is multiplied by a factor of 0.1.

Equation 2.12.2
Where:
s = Rate of condensation (kg/h)
= Heat emission (W/m) (from Table 10.5.1)
L = Effective length of pipe, allowing for flanges and fittings (m)
f = Insulation factor. e.g.: 1 for bare pipes, 0.1 for good insulation
hfg = Specific enthalpy of evaporation (kJ/kg)

Equivalent lengths:
Pair of mating flanges 0.5 m
Line size valve 1.0 m

Example 10.5.1
50 m of 100 mm pipe has 8 pairs of flanges and two valves, and carries saturated steam at 7 bar g.
Ambient temperature is 10°C, and the insulation efficiency is given as 0.1

With reference to Table 10.5.1 and the application of Equation 10.5.1: determine the quantity of
steam that will be condensed per hour:

Part 1 - Without insulation.

Part 2 - With the pipe insulated, but the valves and flanges are left without insulation.

Part 3 - Completely insulated.

Equivalent length of fittings:

(8 pairs of flanges @ 0.5 m) + (2 valves @ 1.0 m) = 6.0 m of pipe

Saturated steam at 7 bar g:


Steam temperature = 170°C
Temperature difference (pipe to ambient temperature) = 170°C - 10°C = 160°C
Enthalpy of evaporation (hfg) = 2 048 kJ/kg
Heat loss per metre of 100 mm pipe (from Table 10.5.1) = 999 W/m

Part 1 - Without insulation:

Equation 2.12.2

Part 2 - Pipe insulated, but without insulation on the valves and flanges:
Consider the two elements separately:

Total condensing rate = heat loss from pipes + heat loss from fittings

Total condensing rate = 8.78 kg/h + 10.54 kg/h = 19.32 kg/h

Part 3 - Pipe and fittings insulated:

Top

Relevant UK and International Standards


Symbols have been used to indicate, technically equivalent standards (=), and related standards (�)
respectively.
Table 10.5.2
Top

Вам также может понравиться