Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 25

Vapor Lock in a Gasoline Fuel Supply

System

Questions: Can PIPE-FLO


Professional model vapor lock in a gasoline
fuel supply system?
Answer:  Yes.

Vapor lock occurs when the gasoline’s actual pressure falls below the gasoline’s vapor
pressure point, somewhere along the fuel supply system. When this occurs, the gasoline
changes from a liquid phase to a vapor phase and there are not enough gasoline molecules
reaching the combustion chamber to maintain the proper air/fuel ratio. As a result, the engine
does not run properly.

Characteristics that may affect the gasoline's actual pressure falling below the vapor pressure
point are listed below.

Gasoline Formulation:

 Summer formulation
 Winter formulation

Pressure Drop

 The suction of the pump is too high.


 The pressure drop across the fuel line is too high. The pressure drop across the fuel
line is too high. This could be because the pipeline is too long, has too small of a
diameter, and/or too many fitting or 90° elbows.
 The pressure drop across a fuel filter is too high.

Increase in Gasoline Temperature

 The ambient temperature increases, and thus the gasoline temperature increases.
 The fuel line or fuel pump is too close to the motor and absorber heat from the motor.
This heat transfers into the gasoline and thus the gasoline temperature increase.

Elevation Change

 The fuel line rises up to a high point, then declines or falls back down creating a
siphon effect.

Combinations of the above

 Pressure Drop plus Temperature Increase


 Temperature Increase plus Elevation Change

Today's vehicles have the fuel pump located in the fuel tank. This solves most of the vapor
lock issues by the pump having a flooded suction and being cooled by the gasoline. However,
if the fuel line or fuel filter is too close to the motor (temperature increase) and there is a
siphon effect (elevation change), a vapor lock can still occur.

Modeling a Vapor Lock in PIPE-FLO Professional:

The first step is obtaining the actual design information.

 Tank elevation and gasoline tank level


 Pipe sizes and pipe elevation changes
 Fuel filter pressure drop and elevation
 Fuel pump elevation, flow rate, and discharge pressure
 Fuel injector or carburetor elevation and any pressure requirements
 Any temperature engine hot spot locations, this would change the gasoline
temperature in the fuel line.

Following is sample PIPE-FLO Professional model showing a vapor lock.

This model shows a tank, fuel line, filter, pump and a pressure source. A pressure source was
used to model the fuel injector. PIPE-FLO turns devices red when there is a warning
message. In this case, the message is "fluid changes state" and the message is shown below.
This message is saying the system pressure has gone below the gasoline vapor pressure and it
is changing from a liquid state to a vapor state. The gasoline fluid zone temperature is 213° F
in this model.

Examples of gasoline temperature verses vapor pressure fluid property relationship are shown
below.

Temperature Vapor Pressure


160° F 4.797 psi absolute
200° F 11.29 psi absolute
213° F 14.58 psi absolute

One can see how the vapor pressure increases as the temperature increases and the maximum
fluid table temperature is 213°F.

Each of these temperature points can be a separate fluid zone. The higher temperature fluid
zone can then be added to the fuel line closest to a hot spot on the engine for sensitivity
analysis.

Additional sensitivity analysis is shown in the model below. This model has the filter
pressure drop to be 0.2 psi g, the filter at the same elevation as the tank, and the gasoline
temperature being 210° F. Under these conditions, the gas is not changing states.
Newtonian and Non-Newtonian Fluids
by Engineered Software, Inc.

PIPE-FLO® Professional can be used to analyze systems with Newtonian fluids.  The


program does not handle Non-Newtonian fluids.

A fluid is defined as a substance that undergoes continuous deformation when subjected to a


shear stress.  The resistance of the fluid to this deformation is ordinarily measured in terms of
the fluid property called viscosity.
Newtonian fluids have a constant viscosity at a given temperature.  For these fluids, the
viscosity is independent of the shear stress rate of change.  Examples of Newtonian fluids are
water, mineral oil, and gasoline.

A shear diagram for a Newtonian fluid is shown below.  Note that the slope is constant.

Non-Newtonian fluids have a variable viscosity at a constant temperature.  The viscosity for


these fluids varies with the rate of shear of the fluid.  Examples of Non-Newtonian fluids are
ketchup, blood, yogurt, gravy, and quicksand. 

A shear diagram for a typical Non-Newtonian fluid is shown below.  Note that the slope is
not constant.
If you picture a fluid as a series of layers, the resistance to flow arises from the friction
between the fluid layers.  According to Newton, when sliding one layer over another layer
twice as fast, the resisting force is two times as great.  Non-Newtonian fluids do not follow
this model.  For these types of fluids, doubling the speed does not necessarily double the
resistance.  It may more than double the resistance (shear thickening, like corn starch in
water), or it may less than double it (shear thinning, like ketchup). 

Safety Tip:  

If you ever fall into a pit of quicksand, its Non-Newtonian behavior is the reason why you
should wiggle out slowly rather than frantically thrashing about.  If you add stress to the
quicksand, it's viscosity decreases and it acts as a liquid which is why you originally
sink.  However, after its initial liquefaction, quicksand's apparent viscosity increases
substantially due to the formation of sand sediment (caused by the removal of water from the
quicksand around your legs).  In order to loosen the sand sediment, you must re-introduce
water into the mixture by wiggling your legs to create areas for water to come into contact
with the sand sediment.  This reduces the viscosity of the sand sediment around your legs
allowing you to escape.

Compressibility and Standard Flow Rates


Q.  When I convert the standard
flow rate of a gas to mass or volumetric, do
I need to account for compressibility?
A.  Yes and no. Oftentimes the compressibility factor (Z) is neglected (or really assumed to
be unity) when performing such conversions, but you must understand the assumption that
warrants its omission. Converting standard flows to mass or volumetric flows is typically
done using rearrangements of the ideal gas equation. Therein lies the assumption; can the gas
you are working with be approximated as an ideal gas under the operating temperature and
pressure of the system? If so, using the ideal gas equation is fine, otherwise Z must be
determined and included in the gas density calculation. The following equation gives a basic
relation between mass, standard volumetric, and actual volumetric flow rates:

w=qAρA=qSρS

Where,
w = mass flow rate
q = volumetric flow rate
ρ = gas density
subscript A = evaluated at actual conditions
subscript S = evaluated at standard reference conditions

The density of a gas can be calculated as:

ρ=PMrZRT

where,
P = pressure
Mr = relative molecular mass
Z = compressibility factor
R = universal gas constant
T = temperature

This is essentially the ideal gas equation, but with the inclusion of the compressibility factor
to account for real gas behavior.  If we assume an ideal gas then Z = 1 and the term falls out.
Combining the two equations above so that we may readily make whatever conversions
needed gives:

w=qAPAMrZARTA=qSPSMrZSRTS

This can be rearranged to give a very common conversion equation to calculate actual
volumetric flow rate when given a standard flow rate.

qA=qS(PSPA)(TATS)(ZAZS)

Oftentimes when this equation is encountered, the compressibility factors (ZS and ZA) are
taken to be 1, thereby assuming ideal gas behavior at both the actual and reference
conditions.  It is valid to assume the gas behaves ideally at the standard reference conditions
(i.e., ZS = 1).  Despite the variety of standard reference conditions used, which tend to vary
by country, industry, standards organization, and maybe lunar phase – it turns out they’re not
all that “standard”, they typically fall in the range of 14.50 to 14.73 psi a (100.0 to 101.6 kPa)
and 32 to 68°F (0 to 20°C).  Under these temperatures and pressures most gases can be
approximated as ideal.

It is the actual conditions under which the gas is flowing where you need to be careful about
assuming ideal gas behavior.  That is, you cannot always assume that ZA = 1, even though a
quick search for conversion equations will turn up many that make this assumption without
acknowledging it.  Many gas systems do operate under conditions in which the ideal gas
assumption can be made, but it is always worth a quick check to ensure that is the case.  To
determine whether a gas at a particular state deviates from ideal behavior, Z can be calculated
from a cubic equation of state like Peng-Robinson or read from a generalized compressibility
chart.  You might also find software applications or online calculators that will calculate gas
density, but once again you need to be aware of the assumptions behind the calculation. 

What Are The Pump Affinity Laws?


Engineers specifying centrifugal pumps for their application are typically most concerned
with the three primary performance characteristics: flow, head, and power. These three
characteristics all vary in orderly fashion with changes in impeller speed, following very
simple equations. These equations are known as the pump affinity laws and allow for the
prediction of pump performance at varying speeds. Given flow, head, and power at one speed
for a specific pump, one can determine new values for those parameters at a different speed. 

The first pump affinity law deals with flow, or pump capacity. The flow varies in direct
proportion with the impeller speed. If the impeller speed for a particular pump is increased by
10%, the flow from that pump increases by 10%. Likewise, if the impeller speed is reduced
by 20%, the flow is reduced by 20%. The affinity law equation for flow is:

Q1Q2=N1N2
Where N1 = Original speed, Q1 = Original flow, N2 = New speed, Q2 = New flow. 
Rearranging we get

Q2=Q1×N2N1

The second pump affinity law deals with fluid head. The head that a particular pump
generates varies with the square of the proportional speed change. For head, when the
impeller speed is increased by 10%, the head is increased by 21% and so on. The affinity law
equation for head is:

H1H2=(N1N2)2

Where N1 = Original speed, H1 = Original head, N2 = New speed, H2 = New head. 


Rearranging we get

H2=H1×(N2N1)2

The third pump affinity law deals with hydraulic power. The power that a particular pump
generates varies with the cube of the proportional speed change, and thus is most impacted by
speed adjustments.  The affinity law equation for power is:

P1P2=(N1N2)3

Where N1 = Original speed, P1 = Original power, N2 = New speed, P2 = New power. 


Rearranging we get

P2=P1×(N2N1)3

Using the equations above, a single point or even an entire pump curve can be adjusted to
find out how the pump in question will perform at a different speed. This is especially helpful
when looking at the possible applicability and benefits of using a Variable Speed Drive.
What is the relationship between pressure
drop and flow rate in a pipeline?
by Jeff Sines, Senior Product Engineer at Engineered Software, Inc.

To understand the relationship between the pressure drop across a pipeline and the flow rate
through that pipeline, we need to go back to one of the most important fundamental laws that
governs the flow of fluid in a pipe: the Conservation of Energy, which for incompressible
liquids, can be expressed using the Bernoulli Equation.

Total Fluid Energy

Daniel Bernoulli, a Swiss mathematician and physicist, theorized that the total energy of a
fluid remains constant along a streamline assuming no work is done on or by the fluid and no
heat is transferred into or out of the fluid. The total energy of the fluid is the sum of the
energy the fluid possesses due to its elevation (elevation head), velocity (velocity head), and
static pressure (pressure head). 

Total Energy=Elevation Head+Velocity Head+Pressure Head

or:

TE=Z+v22g+144Pρ

where (in US Imperial Units):

 TE = total fluid energy (feet of fluid)


 Z = elevation of the fluid measured from a reference plane (feet)
 v = fluid velocity (feet/sec)
 g = gravitational constant (32.2 ft/sec2)
 P = static pressure (psi, or lb/in 2)
 ρ = fluid density (lb/ft3)
 144 = conversion from in2 to ft2

Bernoulli Equation Applied Between Two Points

In reality, the flow of fluid between two points cannot be achieved without a loss of fluid
energy due to friction and changes in momentum. The energy loss, or head loss, is seen as
some heat lost from the fluid, vibration of the piping, or noise generated by the fluid flow.
Head loss is a reduction in the capability of the fluid to do work and will act to reduce the
static pressure of the fluid. 

Between two points, the Bernoulli Equation can be expressed as:


Z1+v212g+144P1ρ=Z2+v222g+144P2ρ+HL

where:

 1 = upstream location
 2 = downstream location
 HL = head loss between the upstream and downstream locations (feet of fluid)

The Bernoulli Equation can now be re-arranged to show that the change in static pressure (dP
= P1 - P2) is due to three components: a change in elevation, a change in velocity,  and the
energy lost to heat, noise, and vibration:

(P1−P2)=ρ144[(Z2−Z1)+(v22−v212g)+HL]

Pressure Change due to Elevation Change

The change in elevation can either be positive or negative. In other words, the upstream
location can be at a lower or higher elevation than the downstream location. If the fluid is
flowing up to a higher elevation, this energy conversion will act to decrease the static
pressure. If the fluid flows down to a lower elevation, the change in elevation head will act to
increase the static pressure.

For example, evaluating just the change in elevation, consider 60 °F water flowing up a hill
from an inlet elevation of 25 ft to an outlet elevation of 75 ft:

(P1−P2)=ρ144(Z2−Z1)=62.4lbft3144in2ft2(75−25ft)=21.7 psi

The positive value of dP indicates that the inlet pressure is greater than the outlet pressure, so
pressure will decrease by 21.7 psi solely due to the change in elevation of the fluid as it flows
uphill. Conversely, if the fluid is flowing down hill from an elevation of 75 ft to 25 ft, the
result would be negative and there will be a 21.7 psi gain in pressure and the outlet pressure
will be greater than the inlet pressure.

Pressure Change due to Velocity Change

Fluid velocity will change if the internal flow area changes. For example, if the pipe size is
reduced, the velocity will increase and act to decrease the static pressure. If the flow area
increases through an expansion or diffuser, the velocity will decrease and result in an increase
in the static pressure. If the pipe diameter is constant, the velocity will be constant and there
will be no change in pressure due to a change in velocity.
As an example, if an expansion fitting increases a 4 inch schedule 40 pipe to a 6 inch
schedule 40 pipe, the inside diameter increases from 4.026" to 6.065". If the flow rate
through the expansion is 368 gpm, the velocity goes from 9.27 ft/sec to 4.09 ft/sec. The
change in static pressure across the expansion due to the change in velocity is:

(P1−P2)=ρ144(v22−v212g)=62.4lbft3144in2ft2((4.09 ft/sec)2−(9.27 ft/s

ec)22(32.2ft/sec2))=−0.466 psi

The negative value of dP indicates that the outlet pressure will be greater than the inlet
pressure. In other words, pressure has increased by almost 0.5 psi from inlet to outlet solely
due to the conversion of velocity head to pressure head.

Pressure Change due to Head Loss

Since head loss is a reduction in the total energy of the fluid, it represents a reduction in the
capability of the fluid to do work. Head loss does not reduce the fluid velocity (consider a
constant diameter pipe with a constant mass flow rate), and it will not be effect the elevation
head of the fluid (consider a horizontal pipe with no elevation change from inlet to
outlet). Therefore, head loss will always act to reduce the pressure head, or static pressure, of
the fluid.

There are several ways to calculate the amount of energy lost due to fluid flow through a
pipe. The two most common methods are the Darcy-Weisbach equation and the Hazen-
Williams equation. The root form of the Darcy equation can be found in the Crane TP-410,
among other industrial references, and is:

HL=fLDv22g

where:

 HL = head loss (feet)


 f = Darcy friction factor (dimensionless)
 L = pipe length (feet)
 D = pipe inside diameter (feet)
 v = fluid velocity (ft/sec)
 g = gravitational constant (32.2 ft/sec2)

The Darcy friction factor, f, takes into account the pipe roughness, diameter, fluid viscosity,
density, and velocity by first calculating the Reynolds Number and Relative Roughness. Once
calculated, the head loss can then be converted to a change in pressure using:

dP=ρHL144
There are other forms of the Darcy equation in the Crane TP-410 that use variables with
different units and a numerical constant that combines all the unit conversions together. For
example:

dP=2.161∗10−4(fLρQ2d5)

where:

 Q = flow rate (gpm)


 d = pipe diameter (inches)

The graph below shows the resulting pressure drop for water at 60 F over a range of flow
rates for a 100 foot long pipe for both 4 inch and 6 inch schedule 40 piping.

Summary

To determine the total change in the static pressure of a fluid as it flows along a pipeline, all
three components of the Bernoulli Equation must be considered individually and added
together. A change in elevation may cause the pressure to decrease, a change in velocity may
cause it to increase, and the head loss may cause it to decrease. The net effect will depend on
the relative magnitudes of each change.

It is possible that the static pressure of the fluid actually increases from inlet to outlet if the
change in elevation or velocity results in an increase of pressure greater than the decrease that
results due to head loss.

The old saying that "fluid always flows from high pressure to low pressure" is not quite
accurate. The more accurate way to state this is that "fluid always flows from a region of
higher total energy to a region of lower total energy".
What is the Relationship Between the Flow
Coefficient (Cv) and Resistance Coefficient
(K)?
by Jeff Sines, Senior Product Engineer at Engineered Software, Inc.

Accuracy of hydraulic calculations is critical for the proper design, operation, and
determination of cost for many types of piping systems in residential, commercial, and
industrial applications. It is crucial that the engineer understand and apply the correct formula
to prevent costly mistakes in the sizing and selection of equipment, operating within safety
limits, and avoiding unnecessary modifications later in the process.  One aspect that leads to
mistakes is the misuse of coefficients  that characterize the hydraulic performance of devices
that have a fluid flowing through them.

For example,  the hydraulic performance of a control valve is characterized by its Flow
Coefficient over its range of travel (Cv in US units, Kv in SI units), whereas the hydraulic
performance of other piping systems devices (such as isolation valves, check valves, tees, and
other fittings) is typically characterized by a dimensionless Resistance Coefficient (K). The
two are inversely related in that the flow coefficient represents how much flow capacity an
obstruction allows, whereas the resistance coefficient represents how much resistance to
flow the obstruction presents.

Both concepts express the Conservation of Energy for a flowing fluid through an


obstruction.

The Resistance Coefficient

The Resistance Coefficient, K, is a dimensionless value given by the equation:

(1) K=fLD

Where

f = Darcy Friction Factor, dimensionless

L / D = Equivalent Length Ratio of the obstruction, dimensionless

L = Length of pipe in feet

D = Inside diameter of the pipe in feet

The Resistance Coefficient is used in a form of the Darcy Equation below:

(2) hL=Kv22g
Where

hL = hydraulic energy lost (head loss) from the fluid due to friction, in ft

v = average fluid velocity, in ft/sec

g = gravitational acceleration, 32.2 ft/sec^2

The head loss, in imperial units, is related to pressure drop, in psi, by the fluid density in
lb/ft3:

(3) hL=144dPρ

Re-arranging the Equation 2 to solve for K and substituting into Equation 3 yields:

(4) K=(2gv2)hL=(2g)(144)(dPρv2)

Conceptually, the Resistance Coefficient represents:

K→(Constant)(Change in Hydraulic Energy of the Fluid (dP

term)Kinetic Energy of the Fluid (ρv2 term))

The Flow Coefficient

In its simplest form, the Flow Coefficient is given by the equation:

(5) Cv=QdP(ρfluidρwater)−−−−−−−−⎷

Where:

Q = volumetric flow rate (in gpm)

dP = pressure drop across the obstruction (in psi)


ρ  fluid = density of fluid (lb/ft3)

ρ water = density of water at standard conditions, 62.37 lb/ft3

This makes gpm/sqrt(psi) the units of the Flow Coefficient, which makes it difficult to see
what the Cv represents conceptually. The equation can be manipulated to bring the concept
into focus.

The volumetric flow rate, Q in gpm, is equal to a unit conversion constant times the flow
area, times the average fluid velocity:

Qgalmin=(7.4805 galft3)(60 secmin)(Aft2)(vftsec)=448.8 Av
(6)
galmin

Where:

A = cross-sectional area of the flow path, in ft2

Squaring both sides of Equation 5 and substituting in Equation 6 for Q gives:

C2v=[QdP(ρfluidρwater)−−−−−−⎷]2=[(448.8Av)2dP]
(7
)
(ρfluidρwater)=[(448.8A)2ρwater](ρv2dP)

Since the area and reference density of water is constant, conceptually the square of the
Cv represents:

C2v→(Constant)(Kinetic Energy of the FluidChange in Hydraulic

Energy of the Fluid)

Relationship Between Cv and K

To determine the relationship between Cv and K, Equation 4 and Equation 7 can be re-
arranged to solve for dP and equated to each other:
Solve Equation 4 for dP:

(8) dP=Kρv2(2g)(144)

Solving Equation 7 for dP:

(9) dP=[(448.8A)2ρwater](ρv2C2v)

Equating Equations 8 and 9 yields:

(10) Kρv2(2g)(144)=ρ(448.8Av)2ρwaterC2v

Simplifying:

(11) K=(2g)(144)(448.8)2A2(62.37)C2v

For circular flow passage with diameter, d in inches:

(12) A=π4D2=π4(d12)2

Where: 

A = pipe flow area in ft2

D = pipe diameter in feet

d = pipe diameter in inches

Finally, substituting Equation 12 into Equation 11:

(13) K=(2)(32.2)(144)(448.8)2[π4(d12)2]2(62.37)C2v
Simplifying the unit conversions into one constant leads to the familiar equation showing the
relationship between Cv and K:

(14) K=890.9d4C2v

Summary

Another example of confusion found with the application of coefficients is seen in spray
nozzle manufacturing which uses a "Nozzle Discharge Coefficient” shown as a K value.
However, a review of their literature shows that they define K = Q/sqrt(dP). This makes the
nozzle discharge coefficient equivalent to the flow coefficient used for control valves. Using
the nomenclature K for the nozzle discharge coefficient confuses the issue even more.

Because engineers view the hydraulic performance of devices differently, mistakes can be
made if the proper equations are not applied. As this paper points out, these coefficients
essentially represent the Conservation of Energy by defining how much hydraulic energy
changes as the fluid flows through an obstruction.

Correctly understanding the concepts and applying methodologies will prevent costly
mistakes in sizing and selection of equipment.  Also the correct methods will ensure
operating safety limits are properly established to avoid unnecessary future modifications. 

Take at look at this article for the Relationship between the Flow Coefficient (Cv) and the
Discharge Coefficient (Cd), commonly used for safety relief valves and flow meters.

What is the relationship between the Flow


Coefficient (Cv) and the Discharge
Coefficient (Cd)?
By Jeff Sines, Senior Product Engineer, Engineered Software, Inc.

Good communication between various groups involved with fluid piping systems is critical
for the proper design, operation, and determination of cost for many systems in residential,
commercial, and industrial applications. It is crucial that the engineer understand and apply
equations correctly to prevent costly mistakes in the sizing and selection of equipment,
operating within safety limits, and avoiding unnecessary modifications later in plant life. One
potential area for costly miscommunication is the use of coefficients for devices that have a
fluid flowing through them. Manufacturers of various equipment use different coefficients
to characterize the hydraulic performance of their devices, and these difference must be
understood when applying them to calculations involving piping systems.
In a previous article, the difference between the Resistance Coefficient (K) and Flow
Coefficient (Cv) was evaluated and a relationship between the two was derived. The Flow
Coefficient (Cv in US units, Kv in SI units) is typically associated with the hydraulic
performance of a control valve, but other devices such as safety relief valves are
characterized by the Discharge Coefficient (Cd, sometimes designated by Kd), which is also
associated with orifices and nozzles. They are not numerically equivalent, so what is the
relationship between the two?

There are various standards in the U.S. and internationally that are used to size and select
control valves and relief valves, most notably the ANSI/ISA-75.01.01 Flow Equations for
Sizing Control Valves (IEC 60534-2-1 equivalent) and the API Standard 520 Part 1, Sizing,
Selection, and Installation of Pressure-relieving Devices in Refineries. These two standards
can be used to derive the relationship between the Flow Coefficient (Cv) and the Discharge
Coefficient (Cd) for relief valves. There are minor differences in the nomenclature used in
each standard, so for the purpose of this article, the nomenclature will be defined for the
equations below along with the engineering units being used.

Control Valve Sizing Equations

When sizing a control valve, the minimum required flow coefficient is calculated based on
the design flow rate and expected pressure drop across the valve, and a valve is selected that
has a flow coefficient greater than the calculated value. Here's the general sizing equation for
control valves for incompressible fluids according to ANSI/ISA-75.01.01 Equation 1, non-
choked turbulent flow:

(1) Cv=QN1ρ1/ρ0dP−−−−−√

Where: 

Q = volumetric flow rate (gpm, m3/hr, or lpm)

dP = pressure drop across the valve (psi, kPa, or bar)

ρ1 = density of the fluid flowing through the valve (lb/ft3 or kg/m3)

ρ0 = density of water flowing through the valve (lb/ft3 or kg/m3)

SG = ρ1 / ρ0 = specific gravity of the fluid (dimensionless)

N1 = constant that depends on the units used for Q and dP (N1 = 1.0 for units of gpm and
psi)

There are other factors that may be included in the sizing equation to account for piping
geometry, high viscosity, or choked flow conditions. Using U.S. units of gpm and psi, the
flow coefficient equation in its simplest form is:
(2) Cv=QSGdP−−−−√

Relief Valve Sizing Equations

When sizing a relief valve, the minimum required effective area is calculated and a relief
valve is selected that has an effective area greater than the calculated value. The sizing
equation for relief valves for liquids using U.S. units according to Equation 28 in the API 520
standard is:

(3) A=Q38KdKwKcKvSGP1−P2−−−−−−−√

Where: 

A = required effective orifice area (in2)

Q = volumetric flow rate (gpm)

SG = specific gravity of the fluid (dimensionless)

P1 = upstream relieving pressure (psig)

P2 = backpressure (psig)

Kd  = rated discharge coefficient (dimensionless) = Cd = (actual flow rate) / (ideal flow
rate)

Kw  = correction factor for backpressure ( = 1 if discharging to atmosphere or if


backpressure is less than 50% of inlet pressure)

Kc  = rupture disc correction factor, if installed ( = 1 if none installed)

Kv  = viscosity correction factor ( = 1 if Re > 105)

38 = all unit conversions compiled into one constant

Assuming no rupture disc is installed, no viscosity correction, and backpressure < 50% inlet
pressure, the API 520 equation (using Cd instead of Kd) boils down to:

(4) A=Q38CdSGdP−−−−√

Rearranging Equation 4 yields:


(5) 38ACd=QSGdP−−−−√

Relationship Between Cv and Cd

The right hand side of Equation 5 is common with the flow coefficient equation, Equation 2
above. Therefore, for liquids:

(6) Cv=38ACd

A similar evaluation can be done for compressible gases and vapors (using Equation 11a in
the ANSI/ISA 75.01.01 standard and Equation 3 in the API standard 520 Part 1, for
example), but the relationship becomes:

(7) Cv=27.66ACd

The next question is: "Why are the constants different?" The answer is that the discharge
coefficient for a given valve is smaller for a liquid than it is for a gas due to the expansion of
the gas as it passes through the valve. For example, one manufacturer shows the discharge
coefficient for one of their valves in liquid service is 0.579, but for gas service is 0.801. The
ratio of the discharge coefficients is 0.801/0.579 = 1.38. The ratio of the constants in the
above equations is 38/27.66 = 1.37, roughly equal.

Deriving the Numerical Constant in Cv to Cd Relationship

The next question a good engineer will ask is: "Where does the constant 38 come from?" The
answer to that requires some unit analysis of the one-dimensional isentropic nozzle flow
energy balance equation, which is given in Appendix B of the API 520 standard.

Using U.S. units for liquid, the mass flow rate per unit area through a nozzle (mass flux, G)
using Equation B.1 and B.6 in the API standard, is:

(8) G=wa=(2)(g)(144)ρdP−−−−−−−−−−−−√

Where:

G = mass flux in lb/sec•ft2

w = theoretical mass flow rate in lb/sec

a = flow area in ft2

g = 32.174 ft/sec2

ρ = fluid density in lb/ft3


dP = pressure drop across the relief, in lb/in2

144 = conversion between in2 and ft2

Disregarding the unit conversion needed for the moment, the mass flow rate is related to the
volumetric flow rate by:

(9) w=ρQ

Therefore, the mass flux is:

(10) G=wa=ρQa=(2)(g)(144)ρdP−−−−−−−−−−−−√

Solving for area (a) and taking the fluid density into the square root:

a=ρQ(2)(g)(144)ρdP−−−−−−−−−−−−√=Qρ2−−√(2)(g)
(1 (144)ρdP−−−−−−−−−−−−√=Qρ(2)(g)(144)ρdP−−−−−−−−−
1)
−−−−√

The density (ρ) in Equation 11 is the fluid density, but the valve sizing equations use the
specific gravity. Specific gravity is:

(12) SG=ρ1ρ0

Where  ρ = density of water at 60 °F = 62.37 lb/ft3.

(13) ρ1=(SG)(ρ0)=(62.37lbft3)(SG)

Taking this relationship into the area equation yields:

a=Q(62.37lbft3)(SG)(2)(g)(144in2ft2)dP−−−−−−−−−−−−−−
(14)
−⎷

Before we throw in all the units, we need the area in square inches, not square feet, so:
A=a(144in2ft2)=(144in2ft2)Q(62.37lbft3)(SG)(2)(g)
(1
5)
(144in2ft2)dP−−−−−−−−−−−−−−−⎷

Now let's put in all the units:

A=∣∣∣144in2ft2∣∣∣Qgalmin∣∣∣min60sec∣∣∣ft37.48055gal62.37lbft3∣∣∣12∣∣∣se
c232.174ft∣∣∣ft2144in2∣∣∣in2dPlb∣∣∣(SG)1∣∣∣−−−−−−−−−−−−−−−−−

−−−−−−−−−−−−−−−−√

(16) A=0.02632QSGdP−−−−√=Q38SGdP−−−−√

The discharge coefficient comes into the equation above because the flow rate, Q, is the
theoretical flow rate assuming incompressible isentropic flow. The discharge coefficient is
the ratio of the actual flow to the theoretical flow:

(17) Cd=QactualQtheoretical

Rearranged:

(18) Qtheoretical=QactualCd

Substituting into the area equation above (Equation 16) yields the relief valve sizing
equation:

(19) A=Qactual38CdSGdP−−−−√

Summary

Over the course of history, the scientific and engineering study involving fluid flow in piping
systems has resulted in developing different coefficients to characterize the hydraulic
performance of various devices that obstruct fluid flow. Because engineers view the
hydraulic performance of devices differently, mistakes can be made if the proper concepts
and equations are not applied correctly. These can be costly mistakes in sizing and selecting
the wrong equipment which can mean the difference between the system having sufficient
pressure relieving capacity or the system rupturing during a high pressure relief incident. 
Each valve has its own flow coefficient. This depends on how the valve has been designed to
let the flow going through the valve. Therefore, the main differences between different flow
coefficients come from the type of valve, and of course the opening position of the valve.
Flow coefficient is important in order to select the best valve for a specific application. If the
valve is going to be most of the time opened, probably there should be selected a valve with
low head loss in order to save energy. Or if it is needed a control valve, the range of
coefficients for the different opening positions of the valve should fit the requirements of the
application.

At same flow rate, higher flow coefficient means lower drop pressure across the valve.

Depending of manufacturer, type of valve, application the flow coefficient can be expressed
in several ways. The coefficient can be non-dimensional or with units if parameters such as
diameter or density are considered inside the coefficient or just in the equation.

Most of valve industry have standardized the flow coefficient (K). It is referenced for water at
a specific temperature, and flow rate and drop pressure units. Same model valve has different
coefficient for each diameter.

Kv is the flow coefficient in metric units. It is defined as the flow rate in cubic meters per
hour [m3/h] of water at a temperature of 16º celsius with a pressure drop across the valve of 1
bar.
Cv is the flow coefficient in imperial units. It is defined as the flow rate in US Gallons per
minute [gpm] of water at a temperature of 60º fahrenheit with a pressure drop across the
valve of 1 psi.

Kv = 0.865 · Cv
Cv = 1,156 · Kv

Needle valve

Search Needle valve manufacturers

It is called needle valve due to the shape of the closure member. It


consists on a threaded stem with a conical end.

Stems with fine threaded have a slow linear movement when they
turn, therefore a great number of turns are needed to have a full flow
section. This makes the needle valve suitable for regulating flow,
with a minimal waste and without cavitation at important differential
pressures.

You might also like