You are on page 1of 4

Choked flow

The basic liquid sizing equation tells us that the liquid flow rate through a control valve is proportional to the square root of pressure drop. This simple
relationship is shown graphically by the green portion of the graph in Figure 1. (Note that the scale of the horizontal axis is the square root of pressure
drop.) This linear relationship does not always hold true. As the pressure drop is increased, the flow reaches a point where it no longer increases. Once
this happens, additional increases in pressure drop across the valve do not result in additional flow, and flow is said to be choked. Here we will call this
limiting or choking pressure drop the Terminal Pressure Drop, pT. (The same thing is also sometimes referred to as the Allowable Pressure Drop,
pallowable, sometimes as the Maximum Pressure drop, pMax, and sometimes as the Critical Pressure Drop, pCrit.)

Figure 1. Liquid flow in a control valve as a function of the pressure drop across the valve.

Let's take a look at what is happening inside the valve to cause this choking of the flow. When the flow stream passes through the vena contracta (the
point at which the cross sectional area of the stream is at a minimum), the flow velocity reaches a maximum. Conservation of energy dictates that since
kinetic energy at the vena contracta has increased to a maximum, potential energy, in the form of static pressure, must decrease to a minimum. (See
Figure2.) Note that in the figure, p is less than pT and flow is not choked.

Figure 2. Velocity and pressure profile in a control valve without choked flow.

As the pressure drop across the valve increases, the flow also increases. Increasing the velocity at the vena contracta and decreasing the pressure at
the vena contracta. If the vena contracta pressure drops below the vapor pressure, vapor bubbles form at the vena contracta. Because vapor takes up
a much larger volume that the liquid, the vapor bubbles fill the vena contracta and any additional lowering of the downstream pressure simply results in
the bubbles getting bigger, but the flow does not increase. It is the formation of these bubbles in the vena contracta that causes the flow to become
choked.

Cavitation

Figure 3 illustrates the choking process along with the cavitation discussed in the next paragraph. Note that in the figure, p is greater than pT and
flow is choked.
Figure 3. Velocity and pressure profile in a control valve with choked flow and cavitation.

As the bubbles move down stream, the cross sectional flow area opens up, the velocity goes down and the pressure goes up. Now we have bubbles
with an internal pressure equal to the vapor pressure surrounded by a higher pressure. The bubbles collapse in on themselves. This combination of
bubble formation and the resulting choked flow, along with the collapse of the bubbles downstream is called CAVITATION. When the bubbles collapse
they make a popping sound. The result is a noise like gravel going through the valve. This noise can be loud enough to be very annoying and even
loud enough to damage the hearing of a person who is exposed to it for long periods. Also, when the bubbles collapse, they create shock waves that
can cause severe damage to the valve. The appearance of cavitation damage is a rough, cinder like, look. (See the picture of a globe valve plug in the
upper right side of Figure 3.) This damage can happen very quickly, sometimes in as little as a few weeks or months. Because cavitation damage
happens so quickly, we try to avoid cavitation at all costs. Very hard materials give some improvement, but usually the improved performance is not
enough to justify the cost.

Flashing

If we continue to decrease the downstream pressure, we reach a point where the pressure downstream of the valve is less than the vapor pressure of
the liquid.

Figure 4. Velocity and pressure profile in a control valve with choked flow and flashing.

Now, instead of collapsing, the bubbles become larger and very soon transition from liquid with bubbles in it to vapor with small drops of water in it.
This is called FLASHING. The appearance of flashing damage is quite different from cavitation damage, and appears as smooth, shiny rivers and
valleys. (See the picture of a globe valve plug in the upper right side of Figure 4.) The damage mechanism is a sand blasting effect. Downstream of the
vena contracta the flow consists of a large volume of vapor with many tiny drops of liquid. Because the volume increases greatly when liquid vaporizes,
the downstream velocity can be several hundred feet per second, and the high velocity liquid droplets can erode away a valve part. The damage
caused by flashing does not usually happen as quickly as that caused by cavitation. The use of hard or erosion resistant materials can usually bring the
damage to within tolerable limits. Trim parts made of the hard stainless steels, such as 17-4 ph, hold up quite well, and 316ss or chrome moly bodies
do much better than carbon steel. The existence of flashing conditions is dictated by the system (p2 is less than pv) and the valve selection neither
causes or prevents flashing. The noise caused by flashing is usually below 85 dBA and to the author's knowledge there is no method for calculating
flashing noise.

The real situation

Figure 1 and the associated discussion of liquid choked flow is the classical discussion, and implies that there is a sudden transition from non-choked
flow to fully choked flow. In reality, at pressure drops approaching, but below the calculated value of pT, there is usually some formation of vapor
bubbles and some degree of cavitation. Figure 5 shows what really happens as flow transitions from non-choked to fully choked flow.

Figure 5. Actual transition between non-choked and choked flow.

It is interesting to note that current control valve sizing methods do not include a method of calculating where the transition from non-choked to fully
choked flow begins and ends. The current ISA and IEC control valve sizing standards only give formulas for calculating the red and green lines in
Figures 1 and 5.

Preventing cavitation and cavitation damage

The value of pT is a function of both the process conditions (p1, the pressure upstream of the valve and pv, the vapor pressure of the liquid) and the
valve's internal geometry represented by the experimentally determined Liquid Pressure Recovery Factor, FL. Typical values of FL are shown in Figure
6. Note that FL is a function of both valve style and the percentage of valve opening. Higher values of FL are associated with valves that have a lower
potential for cavitation, and smaller values of FL are associated with valves that have a greater potential for cavitation.

Figure 6. Typical values of the Liquid Pressure Recovery factor, FL.

There are several methods of increasing the value of pT and thus reducing the potential for cavitation and the associated noise and damage: (1) The
value of p1 can be increased while keeping p the same by moving the control valve to a location further upstream, or to a location at a lower elevation.
(2) The vapor pressure can be decreased by installing the valve where the liquid temperature is lower, such as the cool side of a heat exchanger. (3) A
valve style with a higher value of FL can be selected. It is interesting to note that in general, as the FL increases, so does the price of the valve. There
are special cavitation resistant adaptations of many of the valve styles that have larger values of FL than those shown in Figure 6, yet which retain the
other desirable features of that style.

Because noise and damage can begin before pT reaches pT, there can be considerable risk in simply applying the old rule of not allowing p to
exceed pT. A more reliable method of preventing cavitation damage in control valves, according to one major control valve manufacturer, is to avoid
valve applications where the calculated noise exceeds limits based on a broad range of application experience. This method works because both the
noise and the damage are caused by the same thing, the collapse of vapor bubbles. This manufacturer's experience has shown that for valves 3
inches and smaller in nominal size, cavitation damage will be kept to a minimum if the calculated Sound Pressure Level, SPL, based on un-insulated
schedule 40 pipe, does not exceed 80 dBA. For 4 inch and 6 inch valves this limit can be increased to 85 dBA, and for valves 8 inches and larger the
limit is 90 dBA.

Some globe valve manufacturers allow application of their valves in services where the pressure drop across the valve, p, is no greater than pT,
while others take a more conservative approach of recommending that the actual p be limited to some manufacturer recommended pressure drop
that is less than pT.

Before actually purchasing a control valve, it is always a good idea to ask for the manufacturer's or the manufacturer's representative's comments
regarding your selection.

You might also like