Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 160

Air-oil cylinders, tanks, and intensifiers

Air-oil systems

Compressed air is suitable for many low-power systems, but air’s compressibility makes
it difficult to control actuators smoothly and accurately. Some low-power systems need
the smooth control, rigidity, or synchronization capabilities normally associated with oil
hydraulics. All of these features are available to low-power circuits by using compressed
air for power and oil for control. Purchased or specially built air-oil circuits give smooth
control when the power requirement is low.

Attached oil-control cylinders

Some manufacturers offer attached oil-filled cylinders to control speed and/or


position, Figure 17-1. These units usually work in one direction of travel in a meter-out
circuit. They operate such things as drill feeds or other actions that may try to pull the
cylinder out. (They also can be used with hydraulic cylinders at higher forces.)

Fig. 17-1. Hydraulically controlled air


cylinder – set up for fast advance, controlled feed stroke, and fast retraction

Most manufacturers offer units with valves in the oil line that can stop flow and/or
bypass the speed control. The stop control allows an air cylinder to be stopped
reasonablyaccurately with very good repeatability. The bypass control makes it possible
to have fast and controlled speeds as the cylinder advances.
The cross-sectional view in Figure 17-1 shows an air cylinder that advances rapidly
with airflow controls until its attaching bracket comes in contact with the fast-advance
stroke-length adjustment. At this point, air cylinder movement is retarded and
controlled by the oil speed-control cylinder as oil flows through a flow control. The air
cylinder cannot move any faster than the oil flow allows during this part of the cycle. A
spring-loaded oil balance cylinder furnishes oil to make up for the differential loss from
rod to cap ends. The air cylinder is controlled by oil flow for the remainder of the cycle.

As the air cylinder retracts and the attaching bracket contacts the rod nut, it pushes the
oil speed-control cylinder back to the start position. A flapper-type 1-way check valve on
the piston with through holes allows fluid to transfer back to the rod end. Excess cap
end fluid is stored in the spring-loaded oil balance cylinder during this part of the cycle.

Some manufacturers offer attached units that are capable of control in both directions of
travel. There also are self-contained air powered cylinders with built in oil cylinders and
reservoirs. Air produces thrust while oil controls speed and/or mid-stroke stop-and-
hold. Some units also have two-speed capabilities. These units look like a standard
cylinder with an oversize rod.

Air-oil tank systems

Another common air-oil system uses low-pressure hydraulic cylinders coupled with air-
oil tanks, Figure 17-2. These tanks hold more than enough oil to stroke the cylinder
one way. An air valve piped to the air-oil tanks introduces compressed air to force oil
from the tanks into the cylinder. Add flow controls and shut-off valves to the oil lines to
give smooth, accurate cylinder control.
Fig. 17-2. Typical air-oil tank
arrangement

When control is only necessary in one direction, the tank on the uncontrolled side can
be omitted. This type of circuit requires very good cylinder seals to prevent air or oil
transfer.

Air-over-oil tanks do not intensify the oil pressure, regardless of the tank’s diameter or
length. The highest possible oil pressure available simply equals the air pressure
supplied.

Several cylinder suppliers offer air-oil tanks that consist of a cylinder tube with two
cylinder end caps held on the tube with tie-rods. A sight glass can be a length of plastic
tubing with air-line fittings attached opposite the air ports. A baffle at the air port keeps
oil from being aerated when air blasts in from the valve. A baffle at the oil port keeps any
vortex formed from sending air to the cylinder. This baffle also keeps returning fluid
from blowing into the air port.

Air-oil tandem cylinders

Tandem cylinders are another approach to using oil for control and air for power.
In Figure 17-3, the single-rod cylinder of the tandem runs on air, while the double-rod
cylinder is filled with oil. Because volume is equal in both ends of the double-rod
cylinder, oil flows from end to end through a flow control and/or shut-off or skip valves
for accurate control of speed and stopping.

Fig. 17-3. Typical air-oil tandem-


cylinder circuit

Two flow controls in opposite directions provide variable speed in both directions. A
bypass flow control around the stop valve would allow for two-speed operation in one
direction. (The second speed must be the slower of the two.)

The skip valve option allows a fast approach with deceleration before work contact. The
deceleration signal would come from a limit switch or limit valve.

The schematic drawing in Figure 17-4 shows tandem cylinders in a synchronizing


circuit. This is a practical way to make two or more air-powered cylinders move in
unison. (Using flow controls to do this produces inaccurate results.) When the air valve
shifts to extend the cylinders they must move at the same time. This is because the
trapped hydraulic oil in the hydraulic cylinders must transfer from the top side of one
cylinder to the bottom side of the other one. If one cylinder stops they both must stop at
the same time.
Fig. 17-4. Circuit to synchronize air-oil
tandem cylinders

Note that the maximum load capability is equal to the capacity of both cylinders’ thrust.
With the load placed as shown, the left cylinder transfers energy to the right cylinder
through the oil. This gives the right cylinder up to twice as much thrust.

A small make-up tank and check valves replenish any leakage in the plumbing or at the
rod seal. If the unit is subject to heating, a small relief valve may be required to keep
thermal expansion from over-pressuring the oil-filled chambers. A shut-off valve
connecting the transfer lines can re-synchronize the cylinders if the piston seals allow
fluid to bypass and the platen gets out of level. Re-synchronization can be handled
automatically with a normally closed, 2-way spool valve and limit switches.

(For other air-oil circuits, see the author’s upcoming e-book, "Fluid Power Circuits
Explained.")

Some precautions with air-oil circuits

Most air-oil circuits operate at 100 psi or less, so any pressure drop in the circuit can cut
force drastically. If oil lines are undersized, cylinder movement will be very slow. Size
most air-oil circuit oil lines for a velocity of about 2 to 4 fps. This low speed requires
large lines and valves, but is necessary if average travel speed with maximum force is
important.

Another common problem with air-oil circuits is that any air trapped in the oil makes
the cylinder performance spongy. The air’s compressibility makes accurate mid-stroke
stopping and smooth speed control hard to attain. Some arrangement should be
provided to bleed any trapped air from the oil chambers. When using an air-oil tank
system, it is best to mount the tanks higher than the cylinder they feed. All lines between
the cylinder and the tanks should slope up to the tanks. Also, if possible, let the cylinders
make full strokes to purge any air. With dual oil-tank systems, incorporate a means for
equalizing tank levels into the design.

The cylinder seals must be as leak free and low friction as possible. Any leakage past the
seals can cause tank overflow, oil misting, and loss of control.

Intensifiers (or boosters)

In some of the foregoing air-oil circuits, the usual 80- to 100-psi pressure may not be
adequate for some operations. This does not mean a hydraulic pump and all the items
related to it must be used. Several manufacturers make air-oil intensifiers that convert
80- to 100-psi shop air into 500- to 40,000-psi hydraulic pressure -- in small volumes
of fluid.

Single-stroke intensifiers

The simplest intensifier is a single rod-end cylinder with a large piston rod. As explained
in Chapter 15, a cylinder with a 2:1 area ratio rod can have pressure as high as twice
system pressure in the rod end. This type intensifier is only available in ratios up to 2:1
unless special oversize rods are specified.
Another simple intensifier can be made by coupling the rod of a large-bore cylinder to
that of a smaller-bore cylinder with the same stroke, Figure 17-5. Supplying the large
bore cylinder with pressurized air or hydraulic fluid forces the hydraulic fluid out of the
smaller bore. The upper cross-sectional view is typical of two cylinders assembled in the
user’s plant from stock air and/or hydraulic cylinders. The lower cross-sectional view is
a purchased assembly that takes less space and eliminates possible mounting and
alignment problems. The purchased unit is limited to piston ratios that can have the
same size rod in both cylinders.

Fig. 17-5. Two types of differential-


cylinder intensifiers

Usually these intensifiers are hydraulic to hydraulic with ratios that are less than 5:1
ratio. Later we’ll see a similar design for air-to-air intensifiers with similar ratios. Never
operate these types of intensifiers above the cylinders’ rated pressure. For all intensifier
designs, output pressure is directly related to the area ratio between the driving piston
and the driven piston (or ram).

The cross-sectional view in Figure 17-6 shows typical construction of two types of 25:1
air-oil intensifiers. They consist of 5-in. bore air cylinders with 1-in. rods displacing oil
from high-pressure oil chambers. The upper cross-sectional view is a dual-head
intensifier that requires some sort of blocking valve to isolate its inlet oil from its outlet
oil. This is usually done with a pilot-operated check valve so flow can return when the
actuator reverses.

Fig. 17-6. Ram-type single-stroke intensifiers

The lower cross-sectional view is a triple-head intensifier that has an integral high-
pressure seal to isolate inlet oil from high-pressure oil after the rod moves
approximately 2 in. There is no need for external isolation because oil can flow freely
either way anytime the ram is retracted.

A single-stroke intensifier must be sized to supply enough oil to make the working
cylinder perform its work before the air piston bottoms out. It is good practice to size the
intensifier for 10 to 15% more fluid than required. Avoid long fluid conductors if
possible because the oil’s compressibility and stretching hose can use up the small-
volume safety output quickly.

The circuit in Figure 17-7 shows a typical high-pressure air-oil circuit using the
components described so far. This could be a press operation that requires a 10-in. total
stroke. The stroke concludes with a 0.25-in. high-pressure stroke that generates 25 tons
of force. Based on a maximum pressure of 2000 psi, a cylinder with a 6-in. bore is
needed to produce the required force. The piston area of a 6-in.-bore cylinder is 28.274
in.2, so the 0.25-in. work stroke produces a volume of 7.07 in.3 of high-pressure oil.
Using a standard 5-in. intensifier with a 1-in. ram, this requires

Fig. 17-7. Typical high-


pressure air-oil circuit

(7.07) (110%) / (0.7854) = 9.9-in. stroke plus 2 in. more for passing the high-pressure
seal for a total stroke of 12 in. The volume of the 6-in. bore X 10-in. stroke high-pressure
hydraulic cylinder is (28.275 in.2) X (10 in.) or 283 in.3, so the air-oil tanks should have
6-in. bores and be 12-in. long.

The cycle starts when the solenoid on the 4-way directional control valve is energized to
send air to the left-hand air-oil tank. Simultaneously, the valve exhausts air from the
right-hand air-oil tank. Oil at air pressure is pushed through the triple-head intensifier
to the high-pressure hydraulic cylinder. The cylinder advances rapidly at low force until
it contacts the work.

At work contact, pressure builds in the left-hand air-oil tank and in the pilot line to the
4-way sequence valve. With supply-air pressure at 80 psi and the sequence valve set for
65 to 75 psi, the valve shifts and cycles the intensifier. As the intensifier extends, after it
travels approximately 2 in., it passes through the high-pressure seal to block low-
pressure oil and force high-pressure oil into the cylinder. Pressure in the work cylinder
can now go as high as 2000 psi to produce the required 25 tons of force.

When the solenoid on the 4-way directional control valve de-energizes, air exhausts
from the left-hand air-oil tank and from the 4-way sequence-valve pilot. The sequence
valve shifts to its original position and the triple-head intensifier retracts. Air also is
directed to the right-hand air-oil tank to pressurize it for the retract stroke of the high-
pressure hydraulic cylinder. After the intensifier retracts past the high-pressure seal, the
work cylinder can retract quickly to end the cycle. Note: only 80 psi acting on the area of
the work cylinder develops retraction force. While as much as 25 tons of force was
generated during the short extension stroke, only 1869 lb are generated during
retraction.

The intensifier could be cycled by other means -- such as a limit switch or a pressure
switch and solenoid valve combination. It could even be operated manually.

Also note: any of the above units could be cycled with hydraulic oil as the driving force.
Usually such hydraulic-to-hydraulic intensifiers are only between 2:1 and 5:1 because
the input pressure can be much higher than typical compressed air.

Reciprocating intensifiers

For higher volumes of intensified fluid, several manufacturers make reciprocating units.
The cross-sectional view and circuit in Figure 17-8 show a typical single-ram
intensifier that uses compressed air for power and pumps oil in the high-pressure side.
These units often are supplied in a ready-to-run condition as pictured. They may cycle
as soon as air is supplied or they may require an external signal to start. Most
reciprocating units supply less that 3-gpm maximum at low pressure and slow to a stop
at maximum pressure.
Fig. 17-8. Reciprocating air-to-
hydraulic intensifier

To produce higher pressures, some units incorporate more than one air cylinder in
series to raise the intensification ratio. These units also come with pressure chambers
and rams on both ends to provide a greater volume of high-pressure oil.

Some manufacturers build reciprocating hydraulic-to-hydraulic intensifiers with ratios


as high as 20:1 to generate pressures up to 12,000 psi. These units supply small volumes
of high-pressure oil from low- to high-pressure input fluids.

Special air-oil units

Several companies manufacture special self-contained air/hydraulic cylinders with


integral tanks and intensifiers that produce low-pressure advance, high-pressure work,
and low-pressure retract strokes. Externally, they appear to be over-length air cylinders,
but they can have output forces as high as 150 tons.

Air-to-air intensifiers

When an application requires a small volume of high-pressure air, try an air-to-air


intensifier instead of a high-pressure compressor. The cross-sectional view and circuit
in Figure 17-9 shows the makeup of a 2:1 intensifier that can almost double output
pressure. Inlet air is delivered to the driving cylinder by a double pilot-operated valve
and to the intensifying cylinder through check valves. As the two pistons stroke to the
right, the full area of the left piston and the annulus area of the right piston are pushing
the right piston’s full area at almost double force. Thus, air exiting the right piston is at
about twice input pressure. The discharged air flows through a check valve and on to the
high-pressure circuit.

Fig. 17-9. Air-to-air intensifier with 2:1


ratio

When the pistons complete their strokes, the one on the right contacts a small integral
limit valve that sends a signal to the double pilot-operated valve and shifts it to reverse
the pistons’ strokes. The same areas and forces push in this direction but they work
against a smaller intensifying area. The intensifier will continue cycling until pressure at
the pressure-air outlet port reaches full pressure. At that point, the pistons stall and
hold pressure until the downstream pressure drops.

These intensifiers will stroke considerably more slowly at about 80% of their maximum
pressure so it is best if the output air pressure is at least 20% above what is required. A
regulator at the working machine can control the actual working pressure so less air is
wasted.

Intensification ratios and output volumes are functions of piston ratios, bore sizes, and
stroke lengths. Outputs up to 250 psi are standard with most manufacturers. Some offer
higher pressures. Very high-pressure units use hydraulic cylinders to drive gas cylinders
to reach pressures as high as 45,000 psi.

(For more air-oil and intensifier circuit designs, see the author’s upcoming e-book,
"Fluid Power Circuits Explained.")
RELATED

CHAPTER 16: Accumulators

QUIZ on Chapter 16 Table of Contents Answers to Quiz 16

Bud Trinkel | May 19, 2007

Hydro-pneumatic accumulators

Hydraulic accumulators

Accumulators make it possible to store useable volumes of almost non-compressible


hydraulic fluid under pressure. The symbols and simplified cutaway views in Figure 16-1
show several types of accumulators used in industrial applications. They are not
complete representations but they illustrate general working principles.

A 5-gal container completely full of hydraulic oil at 2000 psi will only discharge a few
cubic inches of fluid before the pressure drops to 0 psi. If the same container were filled
half with oil and half with nitrogen gas, it could discharge more than 1 1/2 gallons of
fluid while pressure only dropped 1000 psi. This is the great advantage of hydro-
pneumatic accumulators.

Accumulator types

No separator: Some original accumulators were high-pressure containers with a sight


glass to show fluid level. They were filled approximately half with oil and half with
nitrogen gas -- with no separation barrier between them. Before stopping the pump, a
shut off valve at the accumulator discharge port was closed to prevent fluid and gas from
escaping. This type of accumulator is not used on new circuits today, but there still are
many in service.

Gas-charged bladder: Many accumulators now use a rubber bladder to separate the gas
and liquid. A poppet valve in the discharge port keeps the bladder from extruding when
the pump is off. The original design was the bottom-repair style, shown on the left in
Figure 16-1. It is still offered by most manufacturers. The top-repair style on the right is
now available and makes bladder replacement simple and fast.

Gas-charged piston: The gas-charged piston accumulator has a free-floating piston with
seals to separate the liquid and gas. It operates and performs similarly to the bladder
type, but has some advantages in certain applications. A gas-charged piston
accumulator can cost twice as much as an equal-sized bladder type.

Spring-loaded piston: A spring-loaded piston accumulator is identical to a gas-charged


unit, except that a spring forces the piston against the liquid. Its main advantage is that
there is no gas to leak. A main disadvantage is that this design is not good for high
pressure and large volume.

Weight loaded: All gas-charged accumulators lose pressure as fluid discharges. This is
because the nitrogen gas was compressed by incoming fluid from the pump and the gas
must expand to push fluid out. The weight-loaded accumulator in Figure 16-1 does not
lose pressure until the ram bottoms out. Thus 100% of the fluid is useful at full system
pressure. The major drawback to weight-loaded accumulators is their physical size. They
take up a lot of space and are very heavy if much volume is required. They work well in
central hydraulic systems because there usually is room for them in the power unit area.
However, central hydraulic systems are falling out of favor, so only a few facilities use
weight-loaded accumulators. (Rolling mills are one application where space to place
large items is not a problem.) Note that there is often a long dwell time to fill these
monsters.

Diaphragm accumulators: There are also diaphragm accumulators with resilient or


metal diaphragms. They are used where the stored volume is small.

Fig. 16-1. Cross-


sectional views and symbols for hydraulic accumulators
Why are accumulators used?

To supplement pump flow: The most common use for accumulators is to supplement
pump flow. Some circuits require high-volume flow for a short time and then use little
or no fluid for an extended period. Generally speaking, when half or more of the
machine cycle is not using pump flow, the application is a likely candidate for an
accumulator circuit.

The circuit in Figure 16-2 uses several accumulators to supplement pump flow because
the dwell time is 45 seconds out of the 57.5-second cycle time. This circuit’s 22-gpm
fixed-volume pump operates on pressure during most of the cycle to fill the cylinder and
the accumulators. Without the accumulators, this circuit would require a 100-gpm
pump driven by a 125-hp motor. The first cost of the smaller pump and motor plus the
accumulators is very close to that of the larger pump and motor. However, energy
savings over the life of the machine make the pictured circuit much more economical.
Fig. 16-2. Accumulator circuit that
supplements pump flow

One drawback of using accumulators to supplement pump flow is that the circuit must
operate at a pressure higher than needed to perform the work. In the circuit in Figure
16-2, a minimum of 2000 psi is necessary to perform the work. This means the
accumulators must be filled to a higher pressure so they can supply extra fluid without
dropping below the minimum pressure. This circuit uses 3000-psi maximum pressure
to store enough fluid to cycle the cylinder in the allotted time and still have ample force
to do the work. The flow control in the circuit is necessary to keep the cylinder from
cycling too rapidly. An accumulator discharges fluid at any velocity the lines can handle
at whatever the pressure drop is when a flow path is opened.

The circuit in Figure 16-2 uses a fixed-volume pump and an accumulator unloading-
and-dump valve. The valve forces pump flow to the accumulators when pressure drops
approximately 15% below its maximum set pressure. At set pressure, the unloading
valve opens and all pump flow bypasses to tank at 25- to 50-psi pressure drop. While the
pump is bypassing, a check valve keeps the accumulators from unloading to tank. The
dump valve (which is a high-ratio, pilot-to-close check valve) is held closed by pump idle
pressure until the pump shuts down.

To maintain pressure: Another common application for accumulators is to maintain


pressure in a circuit while the pump is unloaded. This is especially useful when using
fixed-volume pumps on long holding cycles. The laminating-press circuit in Figure 16-3
clamps material and holds it at force for one to five minutes. If the pump were flowing
across the relief valve at high pressure for this length of time, a lot of heat would be
generated, wasting energy. With a pressure-compensated pump, energy loss would be
less, but the system might still overheat in a short time.

Fig. 16-3. Using an accumulator to


maintain pressure and/or make up for leakage

Adding an accumulator, flow control, and pressure switch to the fixed-volume pump
circuit allows the pump to unload when pressure is at or above the pressure switch’s
minimum setting. If leakage at the valve or cylinder seals allows pressure to drop about
5%, the pressure switch shifts the directional control valve to pressurize the cylinder cap
end and build pressure back to maximum. The only time the pump is loaded is when
fluid is required. This circuit will laminate parts continuously and does not need a heat
exchanger. The flow control should be set at a reduced rate so the accumulator does not
dump too rapidly when the directional control valve shifts to retract the platen. Flow to
make up for leakage is minor and does not need a high rate.

The accumulator dump valve in Figure 16-3 is a high-ratio pilot-to-close check valve that
is held closed by the low pressure when the pump is unloaded. It opens to discharge any
stored energy when the pump shuts down.

To absorb shock: Fast-moving hydraulic circuits can produce pressure spikes that cause
shock when flow is stopped abruptly. Accumulators can be installed in such shock-prone
circuits to reduce damaging pressure and flow spikes to an acceptable rate -- or
eliminate them completely. (Accumulators can handle other pressure-spike concerns
with some valve additions for special instances.)

Figure 16-4 depicts an accumulator installed to eliminate the pressure spike caused by
sudden flow blockage. The nitrogen charge in this installation should be 5 to 10% above
the working pressure. This keeps the accumulator out of the circuit except during
pressure spike situations. A bladder-type accumulator works best here because of its fast
response to pressure changes. (Use caution when applying accumulators to shock
situations. It is possible to actually increase shock instead of reducing or eliminating it.)

Fig. 16-4. Using an accumulator to eliminate shock caused


by a sudden flow stoppage

As an emergency power supply: Some hydraulically operated machines may always need
to stop in the open position to keep from damaging product or equipment. When a
power failure shuts the hydraulic pump off and the machine happens to be some
position other than open, there needs to be some way to get it open. An engine-driven
standby pump could fill the bill and in some instances might be the best remedy.
Another option is to use accumulators that are charged before the first cycle and held
that way until the machine shuts down. The stored energy is ready to cycle the machine
to the open position in case of a power failure.

The circuit in Figure 16-5 operates a slide gate on a waste material bin that opens
hydraulically to fill a transfer truck. The circuit is located in a remote area that is prone
to power failure, so it was designed to automatically close the gate in case power went
off.

Fig. 16-5. Using an accumulator as an


emergency power supply

The schematic diagram shows the cylinder at rest with the pump running. When the
unit starts, solenoids C and C2 on the normally open 2-way directional valves are
energized. They stay energized while the pump is on. The first pump flow goes through
the check valve and fills the accumulator with enough fluid to extend the cylinder from
any open position. When electrical power is available, the gate can be opened and closed
to dump waste material into the waiting truck. If a truck is filling and a power failure
occurs, the pump stops and all solenoids de-energize. At this point the accumulator is
ported to the cylinder cap end and fluid in the cylinder rod end has a free path to tank.

Notice the manual drain connected to the line between the check valve and the
accumulator. This drain must be opened before working on the circuit. A placard on the
machine warns maintenance personnel of the potential danger if the accumulator is not
drained. Emergency power supplies are the only accumulator circuit that cannot be
drained automatically in most cases.

Accumulator precautions
 Always arrange some method to drain the accumulator at shut down. (At the end of this section,
several ways to drain an accumulator automatically are shown. Plus, there is always the old
standby, a manual drain.) Never work on a circuit with an accumulator until you are sure it is
depressurized.
 Make sure accumulator flow is restricted to a reasonable rate during operation and shut down
to avoid damage to the machine or piping. Accumulators will discharge fluid at any rate the exit
flow path will allow. Such high flow does not last long, but the damage it causes is done quickly.
 Always isolate the pump from the accumulator with a check valve so fluid cannot back flow into
the pump. Without a check valve, accumulator back flow can drive the pump backward -- and
overspeed it to destruction in some instances.
 Check the accumulator’s pre-charge pressure at installation and at least once a day for the first
week of operation. If there is no noticeable loss of pressure during this time, do the next check a
week later. If all is well then, do a routine check every three to six months thereafter. Whenever
the accumulator pre-charge drops below nominal pressure, the volume of available fluid is
reduced and finally the cycle slows.

One way to check accumulator pre-charge is to turn off the pump, allow the accumulator
to empty all oil back to tank, and then connect the items in a charge kit, Figure 16-6.
First remove the gas-valve cap and install the charge kit gauge, hose, and tee-handle
assembly on the gas valve. Next, turn the tee handle in to open the valve and read gauge
pressure. However, every time this operation is performed there is the chance the valve
will not reseat and gas will start to leak.
Fig. 16-6. Charging an accumulator or checking its pre-charge
pressure with a charge kit

To avoid potential gas leakage, Figure 16-7 illustrates two noninvasive methods to check
pre-charge. Both are fast, simple, and can be done almost anytime without a lengthy
interruption of production. Either of these ways gives a fast reasonably close check
without invading any plumbing. They are not 100% accurate, but will be within ±5% of
the gauge reading -- with almost anyone doing them. The method on the left is the least
accurate -- especially when using a glycerin-filled gauge.

The Pump Just Starting method on the left shows a jump in pressure after the pump
starts then a steady climb to set pressure. This first jump is the pre-charge pressure and
the steady climb is during compression of the gas in the bladder or behind the piston.
The length of time between the first pressure jump and reaching system pressure
depends on the volume of the accumulator and the pump output.

Fig. 16-7. Two non-invasive procedures


for checking accumulator pre-charge pressure
The Pump Shutoff From Full Pressure method is easiest and most accurate, especially if
the accumulator dump valve is manually operated. Fluid can be bled off slowly with a
manual dump so the gauge reaches pre-charge pressure slowly.

With this method the system must be at pressure and the accumulator charged at least
above pre-charge pressure. At system shut down either an automatic or manual drain is
opened and pressure starts to fall. Because the gauge is reading oil pressure and the only
reason there is pressure is because of trapped gas above it, pressure will fall to a point
then suddenly drop to zero. Read the pressure as the gauge suddenly drops to zero to
determine gas pre-charge.

This method is the most accurate but is not precise like a gauge reading, so use it for a
cursory check as often as necessary to see if the gas charge is holding.

Accumulator pre-charge pressure

Normally, gas-charged accumulators are pre-charged to approximately 85% of the


system’s minimum working pressure. This assures that the bladder or piston does not
discharge all the fluid during every cycle. If all fluid is evacuated at high rates, bladders
can get caught in the poppet valves and pistons can be deformed when metal hits metal.

In certain applications, this 85% figure may be low because minimum system pressure is
low. In such a case, use a piston-type accumulator because the piston can move up the
bore almost any distance without damage. A bladder accumulator should not be used
when pre-charge pressure is less than half the maximum pressure. This avoids
compressing the bladder so tightly that rubbing action on itself wears holes in it.

Applying accumulators

Many applications can use any type accumulator with equally satisfactory results.
However, there are some cases where one particular style is more responsive or offers a
longer service life. As mentioned in the previous section, the amount of pre-charge
pressure is one reason for selecting a bladder or piston accumulator.

Weight-loaded accumulators respond to pressure buildup slowly so they do not work


well as shock absorbers. Weight-loaded accumulators will reduce but not stop pressure
spikes. Piston accumulators are not as fast as bladder types at responding to fast
increases to pressure. So in these situations, the best choice is a bladder-type
accumulator.

Some accumulator circuits are installed to dampen high-pressure spikes at the outlet of
piston pumps. A piston accumulator in this application cannot respond quickly enough
to do the job. Also, the short stroking distance of the piston and seals can cause
excessive wear to the bore and seals. A bladder accumulator works best in this type
circuit.

Sizing accumulators

Most accumulator suppliers offer information in their literature about sizing


accumulators for any of the above circuits. Many offer computer programs that only
require the input of system requirements. The program then figures accumulator size
and outputs a part number. One company offers a formula and software for use on the
Internet.

Accumulator dump valves

In all the foregoing accumulator applications (except the one for emergency power
supply), the accumulator fluid was drained automatically at shut down. This is very
important because accumulators store energy that can be a safety hazard and can cause
damage to the machine. Here are examples of different types of accumulator dump
valves and circuits.
Figure 16-8 shows one frequently used circuit. A normally open, solenoid-operated, 2-
way directional control valve is teed into the pump line between the isolation check
valve and the accumulator. The solenoid is wired so that it is energized when the pump
starts and de-energized when the pump stops. An orifice in front of the 2-way valve
controls flow when the accumulator is discharging to prevent damage to the valve. This
arrangement works equally well with fixed-displacement or pressure-compensated
pumps.

Fig. 16-8. Circuit that uses a solenoid-operated valve to


dump an accumulator

A note of caution: Some solenoid valves, even though they are designed for continuous
duty, get very hot when energized for long periods. Such overheating can cause varnish
deposits to form and lock the valve’s internal parts in the closed condition after the
pump shuts down. This means the trapped energy does not get discharged and the
accumulator can cause harm to anyone working on the circuit.

The dump circuit in Figure 16-9 is only for pressure-compensated pumps. A packaged
set of valves isolates the accumulator while the pump is running and automatically
dump it at shut down. The package consists of an isolation check valve, a pilot-to-close
check valve, and a flow-control orifice.
Fig. 16-9. Hydraulically operated circuit that isolates and
dumps an accumulator supplied by a pressure-compensated pump

At pump startup, flow goes to the circuit and the accumulator. Pressure from the pump
outlet shifts the pilot-to-close check valve, blocking flow to tank. When the accumulator
is full, the pump compensates to no flow and the circuit waits for a new cycle. When
pressure drops, the pump comes back on stroke and makes up for flow going to the
circuit. At pump shut down, pilot pressure to the pilot-to-close check valve drops and
the valve shifts to open. Now, stored energy in the accumulator is ported to tank through
the orifice. This circuit is very reliable because it depends on system or pump pressure
to close and/or open valves.

A fixed-volume pump must be ported to tank at very low pressure when its flow is not
doing work. A common circuit for unloading a fixed-volume pump and dumping an
accumulator is shown in Figure 16-10. An internally piloted unloading relief valve with
integral check valve forces all pump flow to the circuit and the accumulator until the
system reaches the set pressure. As the control ball starts to relieve, system pressure
pushes against the unloading piston and forces it off its seat. This takes all pressure off
the top of the relief valve poppet. The pump unloads to tank at 25 to 100 psi until system
pressure drops approximately 15%. After that drop, spring force pushes the unload
piston back and pump flow goes to the circuit again.
Fig. 16-10. Hydraulically operated circuit
that isolates, unloads, and dumps an accumulator supplied by a fixed-displacement pump

The accumulator dump valve blocks fluid from going to tank while the pump is running
and opens to discharge stored energy when the pump shuts down. The accumulator
dump valve is a high ratio (up to 200:1) pilot-to-close check valve that is held shut by
the pump's unloaded or work pressure. With a 200:1 area ratio between the poppet and
the pilot piston, 25-psi pressure at the pilot port will stop as much as 5000 psi at the
poppet shut off. This keeps fluid in the accumulator circuit until the pump is shut down.
Then, all stored pressurized fluid flows to tank quickly and safely. (One supplier offers
the unloading relief valve and the accumulator dump valve in a single body. This
combination simplifies piping while offering the same effect.)

Other accumulator applications

Accumulators are also used for systems where thermal expansion could cause excessive
pressure. Cylinders with blocked ports in a high ambient heat area can go to high
pressure if there is no place for expanding fluid to go.

Another use for accumulators is as a barrier between two different fluids. The pump that
uses hydraulic fluid keeps pressure on a circuit that uses water or another incompatible
medium.
One supplier offers low-pressure accumulators as breathing devices for sealed
reservoirs. This keeps airborne contaminants out of the hydraulic oil as the fluid level
rises and falls.

For more circuits and other information on accumulators, see the author’s upcoming e-
book Fluid Power Circuits Explained.
RELATED

CHAPTER 15: Fluid Power Actuators, part 1

QUIZ on Chapter 15 Table of Contents Answers to Quiz 15

Bud Trinkel | Apr 27, 2007

Cylinders, fluid power motors, and rotary actuators

Fluid power actuators

Fluid power actuators receive fluid from a pump (typically driven by an electric motor).
After the fluid has been pressure, flow, and directionally controlled, the actuator
converts its energy into rotary or linear motion to do useful work. Cylinders account for
more than 90% of the actuators used in fluid power systems for work output. Of the
approximately 10% of actuators that produce rotary output, more than 90% are
hydraulic motors, while the rest are some form of rotary actuator.

Single-acting ram cylinders

The symbols and cutaway views in Figure 15-1show single-acting ram cylinders in push
and pull types. Rams can be as small and simple as a service station lift operated by air
over oil, or as big and complex as a 100,000-ton extrusion press.
Single-acting rams often are mounted vertically up and are weight returned. When a
ram cylinder is mounted vertically down or horizontally, it must have some method of
retracting it to the home position. Figure 15-1 shows one method. Small single-acting
pull rams -- mounted alongside the large working ram -- raise and hold it in the up
position with a counterbalance valve (not shown). A directional valve or a bi-directional
pump directs fluid to the push or pull rams to make them cycle. Another retraction
method uses single-acting push rams that oppose the platen movement from the
opposite side. (For a circuit that uses a large-diameter vertical down acting ram
cylinder, see Figure 10-9.) Small ram cylinders may be returned manually or via a
spring.

Figure 15-1. Single-acting ram cylinders


for push and pull applications

Ram cylinders only have seals where the ram passes through the body. Anytime a ram
cylinder drifts from its stopped position, the cause is valve or pipe leakage if no fluid is
coming out around the ram seal.

As the ram moves, stops and guide protrusions on it keep it aligned and indicate
maximum stroke. Usually on large-area rams, the stops tear off the packing gland and
bushing retainers if the ram is not stopped some other way. Most machines using rams
have other methods to keep them from overstroking. (Some only have warning placards
about problems if the ram is powered beyond certain limits.) The guide protrusions and
bushing align the ram in its housing so it runs true.

Figure 15-2 shows another type of ram cylinder. When there is a need for a long-stroke
actuator with a short retracted length, one option is a telescoping cylinder. Although the
majority of telescoping cylinders are single acting, double-acting models are available.
Most telescoping cylinders stroke slowly and cycle infrequently because their
construction is not robust enough for high production applications.

Figure 15-2. Single- and double-acting


telescoping cylindersTelescoping ram cylinders

The cutaway views and symbols in Figure 15-2 depict typical multi-stage telescoping
cylinders. The one on the left is single acting; the one on the right is double acting.

Single-acting telescoping cylinders are usually mounted vertically with the small ram
up. The cylinder then can be weight returned. This arrangement leaves the large ram
with its ports attached to a stationary machine member.

Double-acting telescoping cylinders can be mounted vertically with the small ram down
or horizontally when required. The best mounting position for any double-acting
telescoping cylinder is with the small ram attached to a stationery machine member so
the ports do not move. Long-stroke double-acting telescoping mounted horizontally
need some sort of carrier to support the center section during extension so they will not
sag and wear out seals and bushings prematurely. Also note: the return area may only be
only 10% of the extend area, so the return force is not capable of doing much work. This
small area also requires very little fluid to give maximum retraction speed without
excessive backpressure at the extend port. Another possible problem with double-acting
telescoping cylinders occurs if the retract port is blocked while the cylinder is trying to
extend. Up to a 10:1 intensification can result and the high pressure may damage the
housing or rams. Installing a safety relief valve at the retract port may be necessary if
this port can be blocked or restricted for any reason.

For all telescoping cylinders, make sure the small ram can do the work required. As a
telescoping cylinder starts to extend, the large ram always moves first at a lower
pressure. When the first and subsequent rams bottom out, pressure and speed increase
due to the decreased ram area. If the small ram produces insufficient force, the unit
stops before making a full stroke.

Several suppliers build double-acting pneumatic telescoping cylinders in small sizes,


with up to three stages.

One manufacturer makes a single-acting telescoping cylinder with internal porting and
matching areas that cause all rams to move in unison as they extend and retract. These
cylinders come in a maximum of three stages because the area staging would make any
more rams into a vastly oversized package. An integral combination check and relief
valve allows the rams to be filled and bled and to stroke fully in case of bypass at the
seals. This design’s main advantage is smooth extension and retraction without the
bumps of a typical telescoping unit.

Piston-and-rod cylinders
The cutaway view and symbol in Figure 15-3 is for a typical industrial-grade tie-rod
cylinder. This cylinder includes all the standard features available from most
manufacturers. The names of the parts are what most fluid power glossaries propose,
while the names in brackets may be in common use.

The cap end and head end seal off the tube ends with tube-end seals. Tie rods hold the
assembly together. The tie rods are tightened to a torque that will resist as much as five
times the cylinder’s rated pressure. Tie-rod construction gives the package some
flexibility or stretch without permanent deflection or damage. The piston provides the
area for fluid to work against. The piston seals stop bypass that would waste energy. The
piston rod transmits the force on the piston to the outside of the envelope and is
attached to the work mechanism. The rod bushing and rod seal keep the rod aligned and
stop fluid leaks to atmosphere. Cap end and rod end cushion plungers block high fluid
flow near the end of stroke to allow smooth, no-shock stopping. Cushion-adjusting
screws make it possible to adjust stopping speed, while cushion-bypass checks let the
piston move rapidly as the cushion plungers are leaving their chambers.

Figure 15-3. Typical industrial-grade


single-rod end tie-rod cylinder

The symbol on the left is the detailed symbol for a hydraulic cylinder with adjustable
cushions on both ends. This cylinder also could be as shown as: non-cushioned,
cushioned rod end only, or cushioned cap-end only. (When the energy triangles at the
ports are blackened, the cylinder is pneumatic.)

The simplified symbol shows less detail but represents the same unit. The 2:1
information over a single rod end cylinder indicates that the rod area is half that of the
piston. (Cylinders with 2:1 area ratio will be discussed later in this chapter.)

Figure 15-4. Single-acting spring-return


and extend cylinders

The spring return and extend cylinders in Figure 15-4 illustrate another method of
moving cylinder pistons and rods for some applications. The cutaway views show typical
construction (using a tie-rod cylinder as the basic unit). Many other designs are
available but essentially use similar parts. Notice that the pistons have mechanical stops
to keep the spring from compressing enough to bottom out. Breather ports for air
operation or connections for tank drains for hydraulic cylinders are commonly found at
the spring end. Most manufacturers indicate that the spring is only capable of returning
the piston and rod. It may not be capable of returning the external load. Springs can be
less than reliable and difficult to monitor -- especially when they are internal. Because
there usually is little savings in hookup or operation, use these cylinders with care.

Tandem cylinders
The tandem cylinder in Figure 15-5 can produce almost twice the force from the same
diameter, but it is a little over twice the length. The two cylinders can be independently
piped or drained to give extra force in one direction only or both directions. The center
heads have guide bushings and seals for both sections so a different fluid can also be
used in either end. (See Chapter 17 to learn how tandem cylinders allow oil to control
speed and air as the power source. Circuits for matched and unmatched tandem
cylinders can be found in the author’s upcoming e-book Fluid Power Circuits
Explained.)

Figure 15-5. Tandem cylinders with attached rod

The tandem cylinder in Figure 15-5 has a common rod for both pistons. The tandem
cylinder in Figure 15-6 has two separate pistons and rods and two different stroke
lengths. This combination can be used to get three positive stops from an air or
hydraulic cylinder with no special valves or controls. The stops are mechanically fixed,
so the stop positions are in the same place every time. However, the stop positions only
work for one situation. A four- or five-way directional valve at each cylinder plus flow
controls are all that is normally required to operate this circuit.

Special consideration must be used in circuit design for the unattached tandem cylinder
in Figure 15-6. If the long-stroke cylinder is not restrained while the short-stroke
cylinder extends, it can overtravel and miss the exact position. This problem is
exaggerated with horizontal or vertical rod-down applications. Meter-out flow controls
or counterbalance valves can eliminate the problem, but could increase cycle time in
some cases.

Figure 15-6. Tandem


cylinders with attached rod for three positive stop positions (left) and cap-to-cap cylinders for
four positive positions (right)

The cap-to-cap mounted cylinders in Figure 15-6 depict another way to use pneumatic
or hydraulic cylinders to obtain positive positioning without special valves or
equipment. Two four- or five-way valves and flow controls usually make this circuit
operate smoothly.
Some designers specify double-rod end cylinders such as those shown in Figure 15-7.
These cylinders cost about twice as much as single-rod cylinders and the design has a
second place for fluid to leak. In most cases the reason for using them can be
accomplished by other methods with equal or better results. If you must use a double-
rod end cylinder, remember to allow space for the extra rod and the safety hazard it can
cause. Also, the rod reduces the area on the working side of the piston, so a larger bore
or higher pressure is necessary in many cases.

Figure 15-7. Double-rod end cylinders

A double-rod end cylinder might be specified so that the force and speed in both
directions is the same when flow and pressure are equal. This may be true, but flow
controls and a reducing valve can accomplish the same result at a reduced cost and in
less space. Another alternative is a regeneration circuit, used when producing the exact
speed and force in both directions is not critical. (Regeneration circuits are covered
extensively in the author’s upcoming e-book Fluid Power Circuits Explained.)

It may appear that double-rod end cylinders reduce rod flexure when the cylinder is
fully extended. The rods in their bushings and the piston in its bore provide snug
bearing points -- but allow some play. As the piston nears the end of stroke, two of the
bearing points get closer together, so lateral movement at the extended end of the rod
can increase. It is supposed that the opposite rod will reduce lateral movement and hold
the attached load closer to a centered position. However, from the cutaway it is obvious
the distance between the piston bearing and the opposite rod bushing almost eliminates
any centering effect of the piston. A better way to reduce lateral movement of the
extended rod is to stop the piston short of full stroke – either by an internal stop tube or
externally by machine members. This arrangement requires a longer cylinder but gives
the desired results at a lower cost.

A main reason for using double-rod end cylinders is to mount limit switches to show
cylinder position. A special bracket opposite the attachment end holds the limit switches
and a doughnut-shaped protrusion on the rod contacts them as the piston strokes. For
the same price (and consuming a lot less space), most cylinder manufacturers offer limit
switches that attach to the head and/or cap and are activated by cushion plungers.
Another signal indicator -- especially for pneumatics – is a Hall-effect switch and a
magnetic piston to activate it.

All of the above cylinder-position indicators have one potential major flaw. If the part
attached to the rod end gets disconnected for any reason, the machine still will cycle
when the cylinder moves...even though the disconnected load may be in the way. If at all
possible, mount limit switches on the machine member so its position is never
misinterpreted.

Non-rotating rod cylinders

The cylinders in Figure 15-8 incorporate some method to keep the piston and rod from
rotating as it strokes. A standard cylinder may try to turn as it extends and retracts,
causing it to unscrew from its workpiece. In some applications, the cylinder is expected
to orient the work piece it is driving and keep it aligned with mating parts. All the
designs in Figure 15-8 attempt to accomplish this non-rotating function in different
ways. At best they can keep the rod from turning but none can perfectly guide it when
outside forces are acting to turn it. This is especially noticeable on long-stroke cylinders.
It is always best to guide the workpiece externally and only use the cylinder to cycle it.
Note that the oval-piston design offers the ability to mount cylinders side by side with
minimum rod center distance between them and still produce ample force.

Figure 15-8. Non-rotating rod


cylindersRodless cylinders

The cylinders in Figure 15-9 take up less space on long-stroke applications because
they only need mounting space slightly longer than their stroke length. Conventional
piston-and-rod cylinders require space more than twice their stroke length -- and can be
difficult to conveniently place on many machines.

The earliest long-stroke design is the cable cylinder – shown at top left in Figure 15-9.
A coated cable fitted with a work piece attachment, wrapped around two pulleys, and
attached to a piston in a bore produces reciprocating motion as fluid -- usually air --
enters and exhausts through the ports. These cylinders are usually 4-in. bore or less, and
may have strokes up to 30 ft (or more in certain configurations). Cushions may be
specified when required. The cable is coated with nylon or Teflon so it can slide through
seals with minimal damage to them. However, the coatings are prone to cracking and
eventually will cut the seals until they leak. (The symbol for the cable cylinder is adapted
from a manufacturer’s catalog because ISO does not show one.)
Figure 15-9. Cylinders for long-
stroke applications

The rodless cylinder, top right in the figure, was introduced in the late 1970s. It is even
more compact than a cable cylinder and avoids the coating wear problem. It consists of a
piston in a bore that has a slot open to atmosphere along its whole length. A seal blocks
air from escaping through the slot while the piston is not present. A second
contamination seal keeps debris from filling the slot. The fluid seal and contamination
seal pass through slots in the piston in slots as it reciprocates. The work piece
attachment (connected to the piston) reciprocates to move machine members as fluid
enters and exhausts the cylinder. Bores up to 2 12 in. and strokes as long as 33 ft are
available from several manufacturers. Cushions may be specified when required.

The band cylinder is an alternative to the cable cylinder. Its smooth steel band passes
through seals instead of a coated cable. The magnetic-drive cylinder uses magnetic
attraction to keep the piston and workpiece attachment connected. It operates at
pressures up to 120 psi and will maintain connection up to 180 psi.

One manufacturer has a modified rodless cylinder with a toothed belt and pulley
arrangement to drive the workpiece attachment. It offers the option of an external
output shaft to which a brake can be fitted to stop and hold position. This output shaft
can also drive an encoder to show work piece position or can connect to another unit for
synchronization. It also could act as a low power rotary actuator. It is available in 1- or 1
12-in. bores and up to 177-in. stroke.
Other types of linear actuators

The actuators in Figure 15-10 depict other ways of producing linear force. The rolling
diaphragm, diaphragm, and bellows actuators are single acting. The rolling diaphragm
is capable of long strokes but not long life at high cycles. The diaphragm is designed only
for short-stroke applications only but can have high force due to large areas. All these
single-acting devices use some internal or external method to retract them. Using
vacuum and weight are two of the methods used for retracting bellows. The double-
acting rolling diaphragm operates like a double-acting piston cylinder but is not
designed for long life at high cycles.

Figure 15-10. Double-acting


actuatorsSizing linear actuators

Because most cylinders are round, the formula for figuring their area that most
remember is: π r2. Another formula for circle area is .7854 D2 which does not require
changing to radius before figuring area. There are also data books that have charts for all
standard size cylinders. To determine maximum cylinder force, multiply the known area
by the operating pressure in psi. The standard formula is written F=PA where F is force,
P is pressure and A is area.
For pneumatic cylinders, the usual working pressure is 80 psi even though most plants
cycle their air compressor between 115 to 125 psi. The reason for the lower working
pressure is because line losses, especially during high flow surges, make it impossible to
maintain compressor output pressure over the entire air supply system.

Operating pressure for hydraulic circuits often falls between 1500 and 2500 psi because
most components are rated at 3000 psi. Lower pressures require larger actuators at
higher flows and over-sizes the entire system. Pressures over 3000 psi often require
special components and are usually relegated to very high force applications.

A factor that is often overlooked when sizing cylinders, especially pneumatic ones, is the
force figured by the aforementioned formula is not available until after reaching full
pressure. A pneumatic circuit does not reach full pressure quickly like a hydraulic circuit
does. For example, when tank pressure is high and tire pressure is low it is easy to
rapidly fill a flat tire from a portable air tank. However, as tire pressure starts
approaching tank pressure, pressure drop decreases and the transfer rate slows to a
crawl as rising tire pressure gets closer to lowering tank pressure. Pressure equalization
can take several minutes, adding to cycle time when the tire is an actuator. The 3 1/4-in.
bore air cylinder in Figure 15-11 will not move the load attached to it because force
required and force available do not allow enough pressure drop to move fluid from inlet
to the piston. The 4-in. bore gives nominal speed while the 5-in. bore would give fast
speed.

Figure 15-11. Air cylinder sizing


To get nominal cycle times, always size air cylinders at least 25% above workforce
whether moving a load or applying force. When cycle time must be fast, oversizing up to
100% could be required. Anything above 100% will quickly lessen the effect and is not
worth the expense.

Even with hydraulic cylinders, pressure buildup may be fast but is not immediate.
Oversizing of 10% is usually enough to achieve reasonably fast cycles in most situations.
Other things that slow hydraulic actuator cycle time and pressure buildup include
trapped air, slow shifting valves or valves with long spool overlap, and using pressure-
compensated pumps without an accumulator.

Typical cylinder seals

The cutaways in Figure 15-12 show the standard ways of sealing pistons and rods from
fluid bypass. O-rings and U-cups are resilient and continuous seals that can stop all fluid
bypass so a cylinder can be stopped and held almost indefinitely. Cast-iron piston rings
do not seal completely but do a good job when system shock or high heat is part of the
operation.

Figure 15-12. Standard options for


sealing pistons and rods from fluid bypass
O-ring seals are inexpensive but do not have long life in most applications. They must be
setup with interference so as they wear they continue sealing. This also adds to
breakaway and running friction. They are bi-directional and should not be used in pairs.
With two O-rings on a piston, fluid can bypass and get trapped between the seals
causing a lot of friction. At pressures above 500 psi, always specify backup rings to keep
the O-rings from extruding and being damaged. Always keep O-rings well-lubricated
because they can stick and roll in their cavity. When they start rolling they wear rapidly.

U-cups are long-life, low-friction seals because their flexible lips can be heavily
preloaded without much increased friction. These seals are pressure energized, which
means pressure inside the “U” forces the seal material tightly against the sealing
surfaces. U-cups usually use backup rings at pressures above 1500 psi to keep them
from extruding.

Material for the resilient seals discussed here is commonly Buna N rubber for
pneumatics and many hydraulic cylinders. When synthetic fluid or high heat is
encountered, resilient seals often are made of Viton. Buna N is good for temperatures up
to 250° F with mineral oil and water glycol. Viton works at temperatures of 400° to 450°
F and any fluid used in air or hydraulic circuits. Other seal materials such as leather,
Neoprene, and Teflon are available and are standard for some suppliers.

Automotive-type cast-iron piston rings in sets of two or four seal well but not
completely. A general rule is that they will leak approximately one cubic inch per minute
per inch of bore per one thousand psi. Different end effects have been designed to
reduce leakage at the expansion joint. Placing the expansion joint 180° from each other
for consecutive rings helps slow leakage.

Cylinder cushions
Cylinders that move at high speed need some sort of deceleration method to keep them
from slamming when they reach end of stroke. Some applications with high speed and
heavy loads may need valves and limit switches to give enough time to bring the load to
a smooth stop. For most cases standard cylinder cushions work well. They are 3/4- to
11/16-in. long and can be in the head or cap end or both. Cushions add cost on most
cylinders so should not be specified when unnecessary.

Hydraulic cylinder cushions

The cutaway in Figure 15-13 shows a cylinder with an oversize rod and standard
cushions on both ends. The cushion-adjusting screw and cushion bypass check valve can
only be installed in the cap end because the cushion plunger on the rod end is so large
there is no room for them. The necessity of oversize rods is a problem for most
manufacturers. It is possible to taper the cushion and get a smooth stop but it also slows
the stroke as the cushion plunger leaves its chamber. The best option to control
deceleration of oversize rod cylinders’ extend stroke is with external valves or
proportional directional controls.

Figure 15-13. Operation of cushioned


cylinders

The other problem with oversize rods and rod-end cushions is the high intensification
pressure present when working pressure is high and/or there is a heavy overrunning
load. Pressure in the rod end can easily reach two-to-four times rated pressure each
time the cylinder extends. High pressure can damage tube end seals and piston seals,
and stretch the cylinder tube past its tensile limits.

The top cutaway in Figure 15-13 shows the cylinder piston moving to end of stroke
with full flow exiting through the cushion chambers almost unrestricted. This means the
cylinder can travel fast until the cushion plungers enter their chambers. The bottom
cutaway is what happens after the cushion plungers have entered the cushion chambers.
Trapped fluid decelerates the piston quickly to a speed set by the cushion-adjusting
screws. Deceleration is sudden on the straight cushion plungers shown here because
hydraulic fluid is almost non-compressible. The piston continues to end-of-stroke at a
preset controlled speed without damage to itself or the machine.

On cylinders with standard and some oversize rod sizes, the rod-end cushion would
function the same as one on the cap end. There is always some pressure intensification
on the rod end but this is normally not a problem. When the cylinder starts to extend
again, the cushion bypass check valve opens to allow fluid to the full piston area so it can
extend quickly at full force. Without the bypass valves, takeoff speed would be as slow as
deceleration on retract.

The cutaways in Figure 15-13 show a way of making cushions work. In actual cylinder
design, cushions are built in many ways but the general function is the same. One
company offers a self compensating non-adjustable cushion option without adjusting
screws or bypass checks. Another supplier offers tapered or tapered slots in the plungers
that give smooth deceleration instead of the sudden slowdown. Tapered cushions only
work for a given pressure, load, and mounting position. They must be figured from
information collected prior to building the cylinder.

Pneumatic cylinder cushions


Pneumatic cylinder cushions have similar designs but operate differently because the
fluid is compressible. When the cushion plunger enters a cushion chamber, the trapped
air starts compressing according to Boyles’ Law discussed in Chapter 1. When the load is
light and the starting trapped pressure is high enough, the cylinder stops smoothly
without slamming. When the load is heavy and/or starting trapped pressure is low, the
cylinder slows but still may bang the end of stroke. This situation requires the addition
of external deceleration in the form of valves and/or shock absorbers. Often the rod
cushion is ineffective due to less area for the trapped fluid to work against. As with
hydraulic cylinders, oversize rods exaggerate the problem.

Most suppliers offer standard cushions on both air and hydraulic cylinders. This option
could smooth the operation of certain heavy loads that need extra deceleration distance.
These longer cushions only work well for a fixed load and speed condition. If the
machine has changing loads and/or speeds a cushion is not the way to go. Using
external shock absorbers or proportional valves makes control easy to setup.

Cylinder rod column strength and buckling

Most cylinders are designed to have ample rod strength regardless of mounting style to
rated pressure and strokes to 20 in. Above 20-in. strokes, standard rod size may be too
small to keep it from bending in compression. Some mounting styles are more prone to
buckling problems and need special attention. When an order is placed for cylinders
with long strokes that operate above certain pressures, cylinder suppliers usually offer
engineering help on design changes and/or rod modifications. Most cylinder catalogs
have formulas and/or charts that show how to figure rod column strength for different
mounting styles. In many cases the life of a cylinder is dependent on cylinder mounting
type and rod size.

Another problem encountered in cylinder design is with bearing loads. A cylinder has
bearing points at the rod and its bushing and the piston in the bore. Neither area is
strong enough to carry external loads such as a machine member or other parts. Always
use the cylinder to push and pull -- not as a guide.

Figure 15-14 shows column strength and bearing load problems on long stroke
cylinders and how to overcome them. The vertical cylinder at top left with the heavy
load does not have a large enough rod to lift it. When pressure is applied the undersized
rod will bend and the load will not move. If the load does move, rod flex soon wears the
bushing and allows leakage. Matching rod size to load and pressure makes a workable
design.

Figure 15-14. Cylinders with greater


than 20-in. strokes

The lower horizontally mounted long-stroke cylinder with a standard rod has two
problems -- column strength and high bearing loads. The weight of the cylinder as the
cylinder extends wears on the bearing points and high force will finally bend the rod
instead of moving the work. The oversize rod on the lower, horizontally maintained
cylinder is sized to handle the load at maximum force but does little to help bearing load
problems.

Adding a stop tube to the piston rod keeps the bearing points farther apart and reduces
wear. Notice the overall package is longer now because the stop tube takes up space
inside the cylinder. Always check out bearing side-loading problems before building the
machine. It can be difficult to fit the longer stroke cylinder or more costly to constantly
change rod bushings and seals.

Cylinder mounting styles

The mounting styles shown in Figure 15-15 depict the standard NFPA-approved ways
to mount cylinders. Up to 25 different companies make cylinders that match these styles
in every dimension. This means no supplier has to be the sole source for any cylinder on
a machine. Starting at top left are the least expensive mounting styles. Tie-rod mounts
are extensions of the tie-rod threaded section. These extensions go through a machine
member with nuts installed and tightened to hold the cylinder in place.
Figure 15-15. NFPA-
approved mounting styles
The standard offerings are designated two tie rods extended both ends, MX4, four both
ends, MX1, four cap end, MX2, four head end, MX3. At best, these mounting styles hold
a cylinder in place but not necessarily in a precise location. They should only be
considered when what they are attached to does not travel a rigid path.

The tapped mount, top right, is another inexpensive mount but may be hard to locate
and hold cylinder position. It can create problems on long-stroke cylinders that need to
stretch when pressure is applied. With both ends tied down, the cylinder is not free to
grow from pressure or external force without binding. They are designated by the MS
heading.

The side and end lug-mounted cylinders are another way to rigidly mount a cylinder but
also can be a problem on long strokes. The lugs are extra long so dowel pins can be
installed after a precise position is determined. It is not good practice to have dowels in
both ends of the cylinder or on opposite corners especially on long strokes. They are
designated by the MS heading.

Other rigid mountings are flange mounts in rectangular and square style. They are
designated by the MF and ME headings. The bolt-on flanges are for light-duty operation
or where the flange is in compression. The extended head and cap flanges are extremely
heavy duty and directly replace the bolt on flanges.

The most versatile mounting styles are pivot mounts. They allow freedom of movement
in one or both planes and keep the cylinder in perfect alignment. The intermediate
trunion is very good on long stroke applications since cylinder weight can be balanced
and reduce bearing loads. Care in choosing cylinder mounting style can add years to its
life and eliminate rod seal leaks for long periods. Interchangeable NFPA tie-rod design
cylinders come in bore sizes of 1 1/2, 2, 2 1/2, 3 1/4, 4, 5, 6, 7, 8, 10, 12, 14, 16, 18, and 20
in. Rod sizes available are 5/8, 3/4, 1, 1 3/8, 1 3/4, 2, 2 1/2, 3, 3 1/2, 4, 4 1/2, 5, 5 1/2, 7,
8, 9, and 10 in. The largest rod diameter for any bore is equal to or slightly under half
the area of the piston. This is commonly known as a 2:1 area ratio cylinder. Oversize
rods give fast return at low force, allow for regeneration of rod flow, and allow for higher
push force capabilities due to greater column strength. Many manufacturers make
cylinders in smaller and larger bores but they are not necessarily interchangeable.

Steel mills often use what are called mill cylinders that have bolted-on heads and caps.
They are special to a machine in most cases with little or no interchangeability. Mobile
equipment uses cylinders with welded heads and caps that are designed to be thrown
away when they fail. However, many repair shops cut them apart and repair them.
Mobile cylinders also have screwed-on heads or special snap ring arrangements in their
assembly.

The cylinder in Figure 15-16 has such a small diameter and is so far from the
directional valve that the lines to and from it hold more fluid than it does. In this
situation fresh filtered fluid never cycles through the cylinder so it overheats and all
wear particles and rod ingestion stay in it until it fails. Failure is always premature and
the amount of contamination present is excessive when it is repaired. Changing the lines
to the dual flow path shown in Figure 15-16 eliminates the contamination problem.
Cylinder operation and longevity is greatly increased because filtered fluid constantly
flushes it.

Figure 15-16. Cylinder flushing circuit


A tee is installed at A and B ports and check valves facing opposite directions are placed
at each end. Separate lines run from the check valves to a tee at each cylinder port. Now
all fluid must go to the cylinder through one line and return through the opposite one.
Fresh fluid is continually being fed to the cylinder while used fluid returns to be cooled
and filtered.

The dual flow path circuit is more important when the cylinder is mounted vertically
with long lines. Contamination in this configuration lays on the piston and seals and can
score cylinder tube and cause premature seal wear when allowed to collect.

Part 2
RELATED

HAPTER 15: Fluid Power Actuators, part 2

QUIZ on Chapter 15 Table of Contents Answers to Quiz 15

Bud Trinkel | Apr 27, 2007

Oversize rod cylinders

Cylinders with oversize rods can end up with dangerously high intensification pressures
under certain conditions. The cutaway in Figure 15-17shows a cylinder mounted
vertically with its rod down. The rod area is approximately half the piston area so the
annulus area around the rod is also approximately half the area of the piston. The left
circuit shows the cylinder extending with a meter-out flow control circuit. For the 5-in.
bore cylinder with a 3 1/2-in. rod and the pump compensator set at 3000 psi, pressure
in the rod end would be approximately 5880 psi as the cylinder approaches the work. If
this is a hydraulic press, middle circuit, with 5000 lb of platen, tooling a load-induced
pressure of 499 psi would bring the rod end pressure to 6379 psi. This much pressure
could damage the cylinder seals, over-pressure the flow control, and exceed the rating of
pipe connections.

Figure 15-17.
Pressure intensification on an oversize rod cylinder

The circuit on the right eliminates all of the above problems but still allows speed
control of the cylinder. A counterbalance valve on the cylinder rod end set at 100 to 150
psi above load-induced pressure would keep the platen from falling while at rest or
while it is approaching the work. A meter-in flow control sets cylinder speed and as fluid
enters it, pressure in the cap end only rises to approximately half the counterbalance
valve setting. The cylinder starts extending when pressure in the rod end reaches
approximately 574 psi, which is well below the rating of all components.

Each of the circuits in Figure 15-17 would control cylinder speed but the
counterbalance circuit is the best choice for the reasons given plus it has a lot less energy
loss.

Another reason for using an oversize rod is for regeneration circuits like the ones
in Figure 15-18. Any single rod cylinder will at least attempt to extend with equal
pressure at both ports and this is called regeneration. Whether it actually extends
depends on the load it must overcome, maximum system pressure available, and what
the rod diameter is. This is because the maximum force during regeneration is pressure
times the area of the rod. The piston is in balance during regeneration and serves no
function during the process.

Figure 15-18. Single-rod cylinder regeneration

The standard rod cylinder in Figure 15-18 would extend at the rate of 12.25 in./sec at a
maximum force of 3142 lb. Even if this is ample force to extend the present load, the
amount of flow in regeneration is excessive. As the cylinder is regenerating forward with
flow of 10 gpm from the pump, there would be 52 gpm coming from the head end for a
total of 62 gpm to enter the cap port. This high flow would cause excessive back pressure
and keep cylinder speed slow because the circuit relief valve would be bypassing at
system pressure.

Another reason using a standard rod cylinder is not good practice is that its retract
speed would only be 2.33 in./sec, so overall cycle time would not increase as much as
first thought.

The above scenario is the prime reason for using 2:1 rod area ratios for regeneration.
The lower cylinder in Figure 15-18 is the same bore but has a 31/2 in. oversize rod.
This rod is not exactly 2:1 area ratio because it uses NFPA standard sizes for
interchangeability. All NFPA cylinders have the largest standard rod that is up to but not
over 2:1 area ratio.

The figures for the 2:1 ratio rod now show a net force on extend of 9621 lb. at a speed of
4 in./sec. During regeneration, flow from the head end is 10.4 gpm with a total flow to
the cylinder of 20.4 gpm. This is a good measure of force and a reasonable flow rate that
usually overcomes work resistance at easy-to-handle flow rates.

Retract speed would be 3.8 in./sec, making extend and retract speeds almost equal.
When the rod diameter is exactly 2:1, extend and retract speed and force are identical --
the same as a double rod-end cylinder. However, getting exactly 2:1 area ratios requires
an odd size bore or rod that may require special seals.

The circuits in Figure 15-19 show some standard regeneration setups used for
particular needs. The full-time regeneration is a replacement for a double rod-end
cylinder circuit. With a 2:1 area ratio rod, it will have identical speed and force in both
directions. Even with standard rod diameter cylinders, force and speed are within 10%
to 12% of the same, which is often satisfactory.

Figure 15-19. 2:1 rod


cylinders in regeneration circuit

The full force at pressure buildup example uses a sequence valve to indicate work
resistance and direct head-end oil to tank. A check valve in the regeneration flow line
allows regeneration flow to the cap end and prevents pump flow to tank during the full
force portion of the cycle. This circuit extends fast until work contacts, no matter the
size of the part.
The full force at limit switch circuit uses a normally closed two-way directional control
valve to send head-end flow to tank when a limit switch is made. This cuts cylinder
speed in half so part contact is less abrupt. This circuit protects tooling, allows more
time for visual inspection of alignment, and can give an operator more time to respond
to unsafe conditions.

The circuits in Figure 15-19 may need a counterbalance valve to retard running-away
conditions when the cylinders are vertically mounted. When this is necessary, the
counterbalance valve must be externally drained to eliminate backpressure in the
pressure adjustment chamber.

Adding a bleed-off flow control to the line between the head-end port and the check
valve and after a counterbalance valve allows cylinder speed reduction when required.
For complete coverage of regeneration circuits see the author’s upcoming e-book Fluid
Power Circuits Explained.

Rotary output actuators

Air and hydraulic motors also are other types of actuators that turn fluid power energy
into rotary output. They are pumps in reverse and can come in as many varieties. Unlike
cylinders that are rated in pounds force thrust, motors produce twisting or torque that
turns a shaft. Motors are rated in torque output usually stated in pound-inch (lb-in.),
pound-foot (lb-ft.) American; or Newton-meters, (N-m) Metric.
Figure 15-20. Formula for figuring torque

The example in Figure 15-20 shows how torque is measured and applied to shafts.
With a 500-lb weight hanging from the end of a 2-ft arm rigidly attached to a shaft, the
shaft would have to resist 1000 lb-ft of torque (500 lb X 2 ft) to keep from turning. This
is only true when the weight is at right angles to the shaft. Changing arm length or the
amount of weight changes the amount of torque. As noted in the example, the metric
figure is in Newton-meters and one Newton-meter equals 8.851 lb-in. and 0.7375 lb-ft.

Why use a fluid motor?

The main reason cited for using air or hydraulic motors is that they have high torque in
a small package. Air drills of 1/2 to 3 hp for many rotation speeds and torqus require
less than one-third of the space a comparable electric motor setup would use. This
means space at the work is less cluttered and/or more units can be applied. Hydraulic
motors are even more compact especially in higher horsepower units. The main reason
for the small size is no reduction gearing is required for hydraulic motors and small
planetary units work well with air motors.

 Other reasons for choosing fluid motors are:


 They have instant or almost instant reversing capabilities. Because these motors are so compact
and have little inertia to overcome, they can reverse rotation quickly without damage. Some
motors act as oscillators that never make a full revolution.
 Variable speed capabilities of fluid motors are simple and result in little change in torque when
it is a low-speed/high-torque hydraulic motor. A change in flow with flow controls or variable
volume pumps requires little sophistication when minor speed changes can be tolerated.
Sophisticated servo controls allow speed changes and can maintain tight control at any speed.
 Overload protection is part of any fluid power system and motors have the same ability. When a
motor circuit meets resistance it cannot overcome, it stalls and holds torque without damage to
the circuit or machine. Fluid motors are capable of stalling for long periods without over heating
or damaging themselves. There is some internal leakage while stalled but this can be minimized
with the right motor selection.
 Another place where fluid motors shine is in wet or explosive atmospheres. These motors have
no electrical input so they pose no threat from sparks or overheating. A hydraulic motor can
operate underwater with bio-degradable fluids without any of the problems electric motors have
in this application.

Pneumatic motors

Air motors are often very inefficient and with air compressor inefficiency added there is
often no more than 20% utilization of input power. This high inefficiency is offset
somewhat in most applications because the air motor only has to run while work is
taking place. This could mean the air motor must start and stop often but this is not a
problem for fluid motors.

Figure 15-21. Vane-type air motor

There are a variety of pneumatic motors in different configurations and with different
attributes. The vane type shown in Figure 15-21 is one of the oldest designs. They are
usually high to very high speed from 1000 to 30,000 rpm for everything from sanders
and grinders to dentist drills. These motors may offer unidirectional or bidirectional
rotation according to their design use.

The cutaway shows vanes in a rotor that is off center in a cam ring. Air at the inlet acts
against the vanes halfway around, forcing the rotor to turn while spent air exhausts
during the other half revolution. Only one vane in this unbalanced design is producing
torque. The remaining vanes on the inlet side have pressure on both sides so they have
no force.

The vane motor in Figure 15-21 is for medium-to-high-speed applications at low


torque. Several manufacturers add reduction gears to give low-speed/high-torque
capabilities.

Figure 15-22. Radial-piston air motor

The radial piston motor in Figure 15-22 does not require reduction gearing for most
low-speed/high-torque applications. Force from half the pistons is driving the
crankshaft to turn it while the remaining pistons are exhausting. Inlet-outlet ports
connected to a rotary valve driven by the crankshaft send 40 to 100 psi air through
cored holes in the body and cylinder bores to and from the pistons. Most radial piston
air motors run at less than 500 rpm due to air energy waste at faster speeds. These
motors can be physically large and take a lot of mounting space. Some manufacturers
make axial or in-line piston motors while a few have experimented with gerotor designs.

Hydraulic motors

Hydraulic motors come in the same variety as pumps. Many are low-speed/high-torque,
some are high-speed/low-torque, and a few are low- or high-speed/high-torque. The
main difference between pumps and motors is that a motor is usually capable of having
either port pressurized.

High-speed motors can reach 3000 rpm continuous to drive fans, lawn mower blades,
and grinders. They usually do not have high torque starting capabilities but most
applications they are used on do not require this feature. Low-speed/high-torque
motors usually have 75% to 90% of their maximum torque to start. They usually operate
at 500 rpm or less. Piston motors of the in-line and bent axis design have high low speed
torque and can run at 1500 to 2500 rpm without losing efficiency.

Hydraulic gear motors

The gear motor shown in Figure 15-23 is one of the oldest designs and is built for high-
speed/low-torque needs. At first it appears fluid entering the lower port pushes against
two teeth to start the gears turning. However, a closer examination shows the left gear
has fluid pushing on opposing teeth as it comes out of mesh and only the right gear has
any twisting action. After one tooth of revolution, the left gear drives while the right gear
is balanced and so on as the motor turns.

Figure 15-23. Gear-on-gear hydraulic


motorHydraulic gerotor motors

The high-speed gerotor motor in Figure 15-24 has similar characteristics to the gear-
on-gear motor just mentioned. This is not a popular design but the gerotor concept with
the idler gear held stationery shown next is made by many manufacturers and holds
more than 50% of the small-to-medium high-torque/low-speed motor market.
Figure 15-24. High-speed gerotor hydraulic motor

The generated rotor hydraulic motor shown in Figure 15-25 is made high-torque/low-
speed by holding the idler gear still and allowing the orbiting gerotor to cycle inside of it.
This change causes the orbiting gerotor to make as many power strokes as it has teeth
for every revolution of the output shaft. The seven-tooth gear shown makes seven power
strokes while the output shaft turns once. A splined drive connection follows the
orbiting gear and transmits the rotary motion to the output shaft.

Generated rotor motors give at or near full torque from about 25 rpm and normally do
not go higher than 250 to 300 rpm. Maximum output torque is directly related to the
width of the gerotor element which may be as narrow as 1/4 in. to 2 in. Pressure ratings
as high as 4000 psi are common from most manufacturers.

Figure 15-25. Low-


speed/high-torque, hydraulic-generated rotor motor
Gerotor motors can have a selector valve that changes the internal rotary valve output to
feed only half the chambers, causing the motor to run at twice the speed and half the
torque. The gerotor design is machined with too close tolerances but must have some
clearance to allow the inner gear to move. This makes it less efficient and as the gears
wear, internal leakage increases. The geroler design has rolling seal points and can be
setup much closer and has less wear for longer life. Most of the geroler types also use a
plate valve which has less leakage and is wear compensating as well.

Hydraulic vane motors

The hydraulic vane motor shown in Figure 15-26 is a very efficient design and works
well for applications at 20 to 3000 rpm. Fluid entering one port pushes against two or
four vanes as they extend in the cavities of the cam ring. Internal porting directs
pressure and return fluid to the working and exhausting vanes. While half the vanes are
being pushed by fluid, the other half are discharging spent oil to tank. The amount of
torque is in direct relationship to the vane area exposed to pressure fluid and the
distance the vanes are from shaft center. Speed is limited to how much displacement
and what size ports the motor has.
Figure 15-26. Vane hydraulic motors

The high-speed/medium-torque design with an elliptical cam ring gets full torque at
approximately 100 rpm and can go as high as 3000 rpm. Because it covers such a broad
speed range it is suitable for many applications where other designs fall short on torque
or speed. The low-speed/high-torque design is designed for approximately 10 to 400
rpm and usually eliminates any need for gear reduction. Using a direct drive eliminates
maintenance problems and makes a smaller package at the work area.

Hydraulic piston motors

The most efficient and versatile hydraulic motors are piston type but they are also the
most expensive. Inline or axial and bent axis types operate smoothly from 10 to 2000
rpm with high torque throughout their speed range. Radial piston types go as low as 1
rpm but usually not higher than 400 rpm. Their main use is very high-torque/low-speed
applications.

The cutaway in Figure 15-27 shows typical construction of inline fixed- and variable-
displacement hydraulic motors. Low-displacement variable motors may be controlled
manually while larger motors need pistons to change displacement. An inline hydraulic
motor shaft rotates as fluid pushes against a piston, forcing its shoe to slide up the
angled swash plate. A small amount of pressure fluid goes through an orifice in the
piston and behind the shoe to keep it from rubbing metal-to-metal while it is producing
torque.

Figure 15-27. Axial or inline piston


motors

With the swash plate at a steep angle, torque is high while speed is usually low. A
shallow swash plate angle gives high speed but less torque. Most manufacturers
recommend a minimum swash plate angle of 15° to 17° for best results. A maximum
angle of 40° to 45° gives good torque and long motor life.

The bent-axis piston motor in Figure 15-28 has the same operating characteristics as
the inline motor but is more rugged and capable of higher operating pressures. Since
there is no sliding action of piston shoes there is less friction and higher torque for a
given energy input. The angle of the cylinder block to the input shaft determines torque
and speed ranges. The greater the angle, the higher the torque and the lower the speed.
Kidney-shaped openings in both inline and bent-axis motors port fluid to and from
pistons as they rotate. Internal leakage is sent to tank through the case drain. Variable-
displacement bent-axis motors are available but not commonplace due to expense and
size.
Figure 15-28. Fixed-volume,
bent-axis hydraulic motor

The radial piston motor shown in Figure 15-29 uses pistons pushing against an
eccentric to produce rotary motion. These motors usually have five or seven pistons with
rods and shoes, with half of them pushing against the eccentric while the other half
return oil to tank. The shoes have high pressure fluid fed to them from the piston
through the rod to keep them from rubbing the eccentric during the power stroke. A
rotary valve attached to the output shaft feeds and exhausts fluid to and from the
pistons as they turn the eccentric.
Some radial piston motors are made with a moveable eccentric that allows different
offset amounts. Usually the offset is full or one-half so a motor with this feature can
have higher speed at lower torque for fast movement. Eccentric-type radial piston
motors are one type of motor that cannot function as a pump without special inlet
considerations.

Figure 15-29. Radial-piston hydraulic


motor

Other radial piston motor designs are similar in action and torque output, but arrange
the pistons in different configurations. One design has the pistons facing in and pushing
outward against a cam-shaped housing. The shaft is connected to the machine and the
housing rotates. It was originally designed as a wheel motor.

Rotary actuators

When rotary output is one or two turns or less, a hydraulic motor could be used but
repeat stopping accuracy could be a problem. There are several designs of rotary
actuators that give rotary output for limited numbers of revolutions (usually under one
revolution).

Figure 15-30 shows one way to achieve rotary action when the motion is 90° or less. A
clevis-mounted cylinder attached to a lever arm that is driving a shaft gives no less than
half the cylinder force times the lever arm length. In the example shown, torque on the
shaft would be approximately 4000 lb-in. when the cylinder starts and finishes and
approximately 6600 lb-in. when the angle between the cylinder and lever arm is 90°.
Often maximum torque is only required at the end of stroke so arrange cylinder
mounting to give the greatest torque at that point. Remember, retract force is less than
extend force which could cause the cylinder to stall when reversed.

Figure 15-30. Clevis-mounted cylinder for


rotary action

The vane-type rotary actuator in Figure 15-31 is a common design for both pneumatic
and hydraulic fluids. The vane has seals around the edges where it contacts the housing
to control leakage. Fluid entering from the left, as shown, pushes the vane away and
forces fluid out the opposite port as it turns the output shaft. A single-vane rotary
actuator is usually limited to 270° rotation or less. A double-vane rotary actuator usually
only turns 90° or less. Torque is equal to the vane area times input pressure times the
radius from the center of the output shaft to halfway across the vane. The symbol shows
a semi-circle to indicate rotary action that is not capable of being continuous in either
direction.

Figure 15-31. Vane-type rotary actuator


Another common rotary actuator is the single-cylinder rack-and-pinion rotary actuator
shown in Figure 15-32. Several companies make these units in single- and dual-
cylinder models with torque outputs up to and above 300,000 lb-ft. They can be
pneumatic- or hydraulic-actuated. This rotary actuator design can turn more than one
revolution because turns are in direct relation to pinion gear size and rack gear length. It
may also be supplied with cushions, and/or stroke limiters for smooth adjustable stops.
Fluid entering ports at the cylinder ends forces the piston to move away and drive the
rack gear against the pinion gear. The pinion gear continues to rotate until the opposite
piston bottoms out. Reversing flow to the pistons reverses output shaft rotation. Torque
is equal to piston area times input pressure times the radius of the pinion gear.

Figure 15-32. Single-cylinder, rack-and-


pinion rotary actuator

The helical gear rotary actuator shown in Figure 15-33 shows another design that can
have more than one revolution of the output shaft. The number of rotations is in direct
relation to gear teeth angle and the non-rotating piston stroke. Fluid entering and
pushing against one side of the non-rotating piston forces it to move and impart the
turning action to the output shaft through the helical gear and nut arrangement. This
rotary actuator design is able to have deceleration built in and may be purchased as a
double-shaft model.
Figure 15-33. Helical gear rotary actuator

The chain-and-sprocket rotary actuator in Figure 15-34 depicts another way of


achieving limited rotary output with more than one turn capability. The number of turns
is in direct relation to the sprocket size and the working piston stroke. It can be used
with pneumatic or hydraulic fluids. Fluid entering the CCW port pushes against the
working piston and the isolating piston at the same pressure. Because the working
piston has more area it will move left while the isolating piston moves right. The
effective thrust is pressure times the area of the working piston minus pressure times
the area of the isolating piston. The result of these calculations times the radius of the
sprocket determines torque.

Figure 15-34. Chain-and-sprocket rotary actuator

There are some other rotary actuator designs, but in most cases, they are variations of
the ones presented here. For circuits using rotary actuators see the authors upcoming e-
book Fluid Power Circuits Explained.
Part 1

HAPTER 14: Sequence Valves and Reducing Valves

QUIZ on Chapter 14 Table of Contents Answers to Quiz 14

Bud Trinkel | Apr 08, 2007

Download this article in .PDF format

This chapter is sponsored by:

Pressure controls (other than relief and unloading valves)

There are some parts of fluid power circuits that need pressure control. (Chapter
9 covered relief and unloading valves that control pressure in pump circuits.) Other
types of pressure controls include sequence valves, counterbalance valves, and reducing
valves. Though the internal works (and the symbols) are similar, these three pressure
controls perform entirely different functions. Sequence valves and counterbalance
valves are normally closed — like relief valves and unloading valves — but they usually
allow bi-directional flow, so they need a bypass check valve in their bodies. Sequence
valves always have an external drain connected directly to tank. Counterbalance valves
are internally drained, except when used in some regeneration circuits.
Reducing valves are normally open and respond to outlet pressure to keep outlet flow
from going above their set pressure. They also can have a bypass check valve. Reducing
valves always have an external drain connected directly to tank. Any backpressure in
this drain line adds to the valve’s spring setting.

Relief valves, unloading valves, sequence valves, counterbalance valves, and reducing
valves are the most difficult to discern on a schematic drawing because their symbols are
so similar. Take extra care when diagnosing a problem to make sure these valves are
correctly identified and their function understood.

Sequence valves

There are times when two or more actuators, operating in a parallel circuit, must move
in sequence. The only positive way to do this is with separate directional control valves
and limit switches or limit valves. This setup assures the first actuator has reached a
specific location before the next operation commences. If there is no safety concern or
possibility of product damage if the first actuator does not complete its cycle before the
second starts, a sequence valve can be a simple way to control the actuators’ actions.

The symbols and cutaways in Figure 14-1 are for hydraulic and pneumatic sequence
valves. The main difference between these valves is that most hydraulic sequence valves
are single purpose and must be used in series with a directional control valve, while
many air sequence valves are pilot-operated directional control valves with an
adjustable spring return. In either case, a preset pressure must be reached before the
valves allow fluid to pass or change flow paths. Many manufacturers offer a direct-acting
internally piloted hydraulic sequence valve like the design shown in Figure 14-1. This
valve can be changed to external pilot in the field if required.
Figure 14-1. Hydraulic and pneumatic sequence valves

Several manufacturers offer pilot-operated sequence valves also. Pilot-operated


sequence valves stay closed to within 50 psi or less of their set pressure. Direct-acting
sequence valves may partially open at pressures that are 100 to150 psi below set
pressure — and thus allow premature actuator creep.

A balanced spool — held in place by an adjustable-force spring — blocks fluid at the


hydraulic sequence valve’s inlet. When pressure at the inlet reaches the spring setting,
pressure in the internal pilot line pushes the spool up to allow enough flow to the outlet
to keep pressure from going higher. Pressure at the inlet never drops below set pressure
when there is flow to the outlet. When outlet pressure exceeds set pressure, the valve
opens fully and pressure at both ports equalizes. Notice that the drain port hooked to
tank must be at no pressure or constant pressure because any pressure in this line adds
to spring setting. (Remember that a sequence valve must always have an external drain.)

A bypass check valve allows reverse flow when the valve is used in a line with bi-
directional flow. In some applications a sequence valve may be externally piloted from
another operation. Most valves can be converted in the field. (The designer should
always change the part number to reflect the conversion.)
Pneumatic sequence valves typically are 5-way directional control valves with adjustable
springs to set their shifting pressure. They are used to start a second operation after the
preceding one finishes. Some older machines have one solenoid valve to start the cycle
and several sequence valves to extend and retract all other actuators. Some precautions:
• A sequence valve shifts on a pressure build-up and may start a second operation
prematurely if an actuator stalls or is stopped for any reason. If personnel safety or
product damage can occur due to an incomplete stroke, don’t use sequence valves.
Instead, use limit switches or limit valves and directional control valves for each
operation sequence.
• When flow controls are required they must be meter-in types. Take the signal to the
sequence valve from the line downstream from the flow control because pressure at this
point will be whatever is required to move the actuator and its load.

The circuit in Figure 14-2 is typical for air-powered machines. Cyl. 1 extends to clamp
a part when an electrical input signal shifts the solenoid pilot-operated valve. As Cyl.
1extends, pressure beyond the meter-in flow control at its cap end becomes as high as
necessary to move the cylinder and its load. With the sequence valve set to shift at 70
psi, Cyl. 2 should not move until Cyl. 1 has extended and securely clamped the part. If
the clamp does not make a full stroke for any reason, the Cyl. 2 extending prematurely
will not damage the part or be unsafe. When the clamp is at 70 psi or higher, the
sequence valve shifts to extend Cyl. 2. Both cylinders can return simultaneously without
causing any problems.
Figure 14-2. Typical pneumatic sequence valve circuit

One great feature of a sequence-operated circuit is it does not matter how far the first
cylinder must move before the next operation takes place. Thick or thin parts are
clamped at the same force before the next operation starts because pressure must build
to the same level to trigger the next sequence.

Cyl. 2 has meter-out flow controls to retard its movement and hold pressure on Cyl.
1during the stamping operation. De-energizing the solenoid pilot-operated valve allows
both cylinders to return home at the same time.

The hydraulic sequence circuit in Figure 14-3 is typical for a machine that must clamp
and hold pressure while a second operation takes place. Sequence valve 1 is set at 550
psi; pressure at clamp Cyl. 1 must be at least 550 psi before punch Cyl. 2 can extend.
While punch Cyl. 2 is extending, pressure in the circuit never drops below 550 psi. If the
punching operation requires more than 550 psi, the pressure in the whole circuit
increases -- up to the relief valve setting.
Figure 14-3. Typical hydraulic sequence valve circuitSequence valve 2 (set at 450 psi)
keeps Cyl. 1 from getting a retract signal until Cyl. 2 has returned and pressure
increases.

A pilot-operated check valve maintains clamp force while the punch cylinder retracts.
The signal to open the pilot-operated check valve comes from the line between Sequence
Valve 2and Cyl. 1, so there is no signal until Cyl. 2 fully retracts. (This circuit is not safe
if pressure buildup comes from some source other than clamp contact or the end of
stroke so that the punch cylinder operates prematurely.)

Figure 14-4. Kick-down sequence valve

Sequence valves often generate a great deal of heat because the first actuator to move
takes higher pressure than the subsequent actuators. This means there is usually a high
pressure drop across a sequence valve that results in wasted energy. In some circuits, a
kick-down sequence valve can reduce the energy loss. The cutaway view and symbol
in Figure 14-4show the inner workings of a kick-down sequence valve to explain how it
controls opening pressure and then unloads it.

Fluid from the inlet flows through the control orifice and up to the adjustable poppet
where it is blocked. The resulting pressure tries to open the poppet while equal pressure
and a light spring acting on the opposite side hold it shut. When pressure increases
enough to unseat the adjustable poppet and more flow starts passing the poppet than
going through the control orifice, the pressure imbalance lets the poppet raise. When the
poppet moves enough to let trapped fluid go through the bypass orifice, pressure on top
of the poppet drops off -- because the bypass orifice is larger than the control orifice. At
this point, the only force acting to hold the poppet shut is spring force and backpressure
at the outlet port. When flow stops, the poppet closes again due to pressure equalization
and spring force on the poppet.

Figure 14-5. Hydraulic circuit with kick-down sequence valves

The circuit in Figure 14-5 is the same as in 14-3 except it incorporates kick-down
sequence valves in place of standard sequence valves. Cyl. 2 will not extend in this
circuit until pressure on Cyl. 1 has reached 750 psi. The difference is when a kick-down
sequence valve opens at its pressure setting, it allows fluid to pass at 50 psi plus
whatever it takes to overcome downstream resistance. This means the whole circuit
from the pump to all actuators is 50 psi plus Cyl. 2’s resistance. The pilot-operated
check valve at Cyl. 1’s cap-end port keeps it pressurized at near full force, while Cyl.
2 extends at low force. Energy waste is very low so heat buildup is minimal. (Other
sequence valve circuits can be found in the e-book Fluid Power Circuits Explained by
the author of this manual, which will be launched in the next few months.)

Counterbalance valves

The fourth and last normally closed pressure control valve found in hydraulic circuits is
the counterbalance valve. Cylinders with external forces — such as weight from a platen,
machine members, or tooling — acting against them will overrun when cycled if oil
flowing out of them is not restricted. A meter-out flow control circuit is one way to
control overrunning loads but it has one main drawback. A flow control’s speed is fixed
except for manual adjustment or when using an infinitely variable proportional type.
Because flow is fixed, the actuator will continue at the same speed – even when working
flow to it increases or decreases. Thus, control is minimal and there could be high
energy waste. (Figure 13-8shows a meter-out flow control circuit for running away
loads.)

Download this article in .PDF format

This chapter is sponsored by:

A counterbalance valve keeps an actuator from running away regardless of flow changes
because it responds to pressure signals, not flow. A counterbalance valve is almost the
same as a sequence valve except it normally does not have an external drain connection.
The cutaways and symbols in Figure 14-6 depict the physical makeup of three different
counterbalance valves and how they are represented on a schematic drawing.

The two cutaways and symbols on the left are spool designs with internal and external
pilots. The valve on the right is a poppet design that is both Internally and externally
piloted. Each valve type has advantages in different circuit arrangements that will be
discussed later. A counterbalance valve usually has a bypass check valve for reverse flow
because its most common use is in controlling actuators with running away or
overrunning loads.

Figure 14-6. Three types of counterbalance valves

An internal pilot-operated counterbalance valve shifts to allow excess fluid to flow to the
outlet when pressure at the inlet increases to the pressure set by the pressure
adjustment. Pressure at the inlet never drops below set pressure when there is flow at
the outlet. Flow from the inlet to the outlet is just enough so that backpressure on the
actuator never drops below set pressure. This means the actuator moves only as fast as it
is supplied and stops when Inlet flow ceases.

Pressure adjustment on the Internal-piloted counterbalance valve is usually made by


first screwing the pressure adjustment all the way in. To assure that the valve is capable
of high enough pressure, start the pump and raise the load a small amount. Then center
the directional valve — which connects the cylinder rod-end port to tank — to see if it
holds. If the load holds, next raise the load in increments — checking for load stop every
few inches. With the load suspended, start reducing set pressure on the counterbalance
valve slowly until the load creeps forward. When the load starts drifting down slowly,
increase pressure until movement stops, then turn the pressure adjustment another
quarter to half turn higher. This method of adjusting usually wastes less energy while it
always stops and holds the load.

The main disadvantage of an internal pilot-operated counterbalance valve is that


backpressure is constant and it holds back even when the actuator needs maximum
force. Another disadvantage is that to maintain optimum performance, an Internal-
piloted counterbalance valve must be readjusted every time the load changes. The
valve’s main advantage is that it produces smooth cylinder action while advancing to the
work.

An external pilot-operated counterbalance valve shifts to allow excess fluid flow to the
outlet when pressure at the opposite cylinder port reaches the pressure set by the
pressure adjustment. Pressure at the inlet never drops below load-induced pressure plus
pressure set on the pressure adjustment when there is flow at the outlet. Flow from inlet
to outlet is just enough that the actuator moves only as fast as it is supplied and stops
when flow to the actuator ceases.

Pressure adjustment on the external pilot-operated counterbalance valve can be made


on a test stand by setting the pressure adjustment at 100 to 200 psi. If pressure must be
set on the machine, set the pressure adjustment higher than 200 psi and lift the load a
small distance to make sure it stops and holds. If it holds, continue to raise the load high
enough to have some time for the next step. Now, power the load down and observe
pump pressure. Pump pressure while lowering the load should not exceed 200 psi.
Continue this action until pump pressure is between 100 and 200 psi while the load is
lowering. This method of adjusting usually wastes less energy while always stopping and
holding the load.

The main disadvantage to an external pilot-operated counterbalance valve is that it may


cause lunging or even stop cylinder action while advancing to the work. The main
advantage is that backpressure is only present when the actuator is advancing to the
work. At work contact, pressure at the actuator inlet increases and forces the
counterbalance valve wide open, thus eliminating all backpressure. Another advantage
is that an external pilot-operated counterbalance valve does not need to be readjusted
when the load changes.

Internal and external pilot-operated counterbalance valves shift when pressure at the
internal pilot area reaches the pressure set on the pressure adjustment and allows excess
flow to go to the outlet. Pressure at the Inlet never drops below set pressure when there
is flow at the outlet. Flow from the inlet to the outlet is just enough that backpressure on
the actuator never drops below set pressure. This means the actuator moves only as fast
as it is supplied and stops when Inlet flow ceases.

Pressure adjustment on an internal and external pilot-operated counterbalance valve is


usually made by first screwing the pressure adjustment all the way in. To assure that the
valve is capable of high enough pressure, start the pump and raise the load a small
amount. Then center the directional valve that has the cylinder rod-end port connected
to tank — to see if it holds. If the load holds, then raise the load in increments —
checking for load stop every few inches. With the load suspended, start reducing set
pressure slowly until the load creeps forward. When the load starts drifting down slowly,
increase pressure until movement stops, then turn the pressure adjustment another
quarter to half turn higher. This method of adjusting usually wastes less energy while
always stopping and holding the load.
An internal and external pilot-operated counterbalance valve lowers loads smoothly and
opens fully when pressure at the actuator inlet increases upon contact with the work.
The valve does need to be readjusted when loads change, but this is a small price to pay
for good control.

Figure 14-7 depicts a vertically oriented cylinder with rod facing down and a load
trying to extend it. To keep the cylinder from running away, the counterbalance valve
must resist the load-induced pressure from the weight. The load-induced pressure can
be calculated and the counterbalance valve could be preset at 100 to 150 psi higher on a
test stand, but pressure adjustment is usually done at the machine (as mentioned
earlier).

Figure 14-7. Internally pilot-operated counterbalance valve circuit

Notice that the directional control valve has ports A and B connected to tank in the
center condition. There is no chance of extra pressure buildup in the pilot line while the
circuit is at rest. If ports A or B were blocked, pressure could build and pilot the
counterbalance valve open, allowing the cylinder to drift.

Energizing solenoid A1 sends pump flow to the cylinder cap end. As pressure builds
there, pressure also increases in the rod end. When pressure at the cylinder rod end
reaches 100 to 150 psi above the load-induced pressure, the cylinder starts to extend as
fast as the pump fills the cap end. When flow increases, cylinder speed increases and
when flow decreases, cylinder speed decreases.

As stated in the counterbalance valve explanation, backpressure at the cylinder rod end
is present during the entire extend stroke. As a result, at work contact cylinder force is
reduced by counterbalance pressure times the cylinder’s rod-end area. The total weight
of the platen and tooling on a press plus the amount of added pressure at the
counterbalance valve cannot be used to do work. Energy is expended to raise the weight
but it is not recouped during the work cycle. Energizing solenoid B1 sends fluid around
the counterbalance valve through the bypass check valve and on to the cylinder rod end
to retract it.

Figure 14-8. Externally pilot-operated counterbalance valve circuit

The circuit in Figure 14-8 shows the same cylinder with an external pilot-operated
counterbalance valve. An externally piloted valve can be set at approximately 100 to 200
psi regardless of load-induced pressure in the cylinder. This is especially convenient in
applications where loads constantly change. It is also the best use of energy because the
counterbalance valve opens fully when the cylinder meets resistance so the weight is
able to do some work. Because backpressure on the cylinder rod end is zero, more force
is available.

Energizing solenoid A1 sends fluid to the cylinder’s cap end to start it extending. As
pressure builds in the cylinder cap end, it pressurizes the external pilot and opens the
counterbalance valve The valve only opens enough to let fluid out when the cap end is at
pilot pressure. If pilot pressure is set too low, the counterbalance valve may quickly open
too far — allowing the cylinder to run away and pilot pressure to drop. At this point, the
counterbalance valve shuts abruptly and the cylinder stops. Almost immediately,
pressure again builds at the cylinder cap end, the counterbalance valve reopens, and the
same scenario repeats until the cylinder meets resistance. A meter-in flow control in the
external pilot line can help, but is very difficult to set. Energizing solenoid B1 sends fluid
around the counterbalance valve through the bypass check valve and on to the cylinder
rod end to retract it.

The internal and external pilot-operated counterbalance valve in Figure 14-


9 incorporates the best features of both valves. The internal pilot provides a smooth
advance stroke at low force, while the external pilot opens the valve fully to eliminate
backpressure from the cylinder rod end when it contacts the workpiece. (Like the
internally piloted valve. this version must be reset at each load change to maintain its
efficiency and keep energy losses low.)
Figure 14-9. Internally and externally pilot-operated counterbalance valve circuit

The symbols in these example circuits show a direct-acting pressure control valve.
Several suppliers offer a pilot-operated version that is more stable and has less pressure
differential between cracking and full flow operation.

The circuits shown here work equally well with hydraulic motors, except that a
counterbalance valve will not stop and hold a running away load on a motor without
creep. All hydraulic motors have internal leakage that increases as the motor wears. The
counterbalance valve may not have any bypass but fluid will slip by the motor parts no
matter what its design.

There are no counterbalance valves for air circuits. Air circuits depend on meter-out
flow controls to keep an actuator from running away. Usually an air circuit uses a 2-
position valve that keeps pressure on the retract side at rest so it stays in place at end of
stroke. When a load must be stopped in mid-stroke, a 3-position valve with cylinder
ports blocked in center is the common method of trying to do this. There also is
available a pilot-operated check valve for air service that gives some control for stopping
and holding a pneumatic cylinder in mid-stroke.

Air line regulators


Download this article in .PDF format

This chapter is sponsored by:

Most plant air systems produce pressures between 90 and 125 psi, while most air
circuits are designed to operate at 75 to 85 psi. Other systems may operate at pressures
as low as 15 to 20 psi. To accommodate these ranges, some method is needed to reduce
the system pressure without wasting energy. A relief valve that would release plant air to
atmosphere and try to lower the whole system it is not a good solution. The air line
regulator shown in Figure 14-10 reduces outlet pressure by shutting off flow when
downstream pressure tries to go above the regulator’s setting. There is very little energy
loss because air merely expands from its elevated pressure to meet the lower pressure
requirement. In other words, an air compressor operating at 120 psi only has to run
about a third as often when regulated or reduced to 40 psi.

Figure 14-10. Air line regulators (or reducing valves)


This points out the main reason why an air line regulator should be set just high enough
to do the job at hand. Without a regulator, not only does it cost more to operate a
machine, but the machine tries to run a repetitive cycle with fluctuating pressure,
therefore different forces and speeds.

The cutaway views and symbols in Figure 14-10 show two common direct-acting air
line regulators. They are normally available in sizes from â…› through 2 in. with pipe
thread. (Larger sizes are built but they usually are pilot-operated from a small direct-
acting regulator.) Air flows freely from inlet to outlet until the outlet pressure reaches
the set pressure. The adjustable spring holds the shut-off valve off its seat by extending
the diaphragm during free flow. As pressure at the outlet continues to build, it passes
through the pilot passage to the underside of the diaphragm. At set pressure, the
diaphragm pushes the adjustable spring back, allowing the shut-off valve to seat. The
light spring pushes the shut-off valve closed. Pressure at the outlet now is stable at its
reduced setting — as long as the inlet pressure is equal to or higher than the outlet. Any
pressure drop at the outlet reduces pressure under the diaphragm and the adjustable
spring again pushes the shut-off valve open to let more air in.

If there is a possibility of the reduced pressure line seeing excess pressure for any
reason, use the relieving-type regulator shown on the right in Figure 14-10. This valve
closes a hole through the diaphragm’s center section with the shut-off valve’s stem. After
reaching set pressure, the shut-off valve cannot move up. Any extra pressure buildup
under the diaphragm raises its center section off the shutoff valve’s stem and allows air
to flow to atmosphere through the vent hole. This feature should not be used as a relief
valve function where pressure increases during every cycle -- it is only for occasional
overpressure situations.

Every pneumatically powered machine should have a regulator set for the lowest
pressure that will produce good products. Costly overpressure should be eliminated in
every case. Use an air line regulator anytime a job can be done at a pressure lower than
plant air supply.

Another application for air line regulators that can save compressor output is reducing
pressure on the return stroke of actuators that can use low power to retract. Many
cylinders need high force to extend and do work, but the retract portion of the cycle
needs very low force. An air line regulator positioned as shown in Figure 14-11 can save
air during part of every cycle on many cylinder operations in most circuits.

Figure 14-11. Pneumatic circuit with air-saving regulator

A 5-way spool valve, piped with a dual-pressure inlet as shown, can give normal cycle
time while conserving plant compressed air. Return pressure is set on the regulator
supplying the cylinder rod end at the lowest possible pressure that maintains cycle
integrity. A reduction as small as 20 psi below working pressure can pay for the
regulator in a short time. Shifting the 5-way valve starts the cylinder extending. (There
will be a brief lunge as the lower-pressure air in the rod end compresses to hold back
against the higher pressure in the cap end.) To control cycle time, adjust cylinder speed
with the rod-end meter-out flow control. When the 5-way valve shifts again to return the
cylinder, the meter-out flow control on the cap end must be adjusted for a faster rate
because return power is limited.

Pressure-reducing and reducing//relieving valves

Download this article in .PDF format

This chapter is sponsored by:

There are times in multi-actuator hydraulic circuits when system pressure is too high for
some actuators while others need maximum force. One suggested remedy is the circuit
in Figure 14-12. Cylinder 1 needs 2000 psi to maintain force, while Cylinder 2 can
damage the product when pressure exceeds 800 psI. Adding Relief Valve 2 (set at 800
psi) takes care of Cylinder 2’s overpressure, but limits the entire circuit to 800
psi. Pressure in a circuit with more than one relief valve will never be higher than the
setting of the lowest valve. The correct way to have two or more pressures in a single
circuit is to incorporate reducing valves. (Figure 14-14 diagrams a circuit using a
reducing valve to give two pressures.)
Figure 14-12. Dual-pressure hydraulic circuit with two relief valves

The cutaway and symbol in Figure 14-13 depicts a pilot-operated reducing valve that
allows flow from the inlet to the reduced-pressure outlet until pressure reaches the
setting on the direct-acting relief valve in the pilot section. Unlike the other four
pressure controls (relief, unloading, sequence, and counterbalance valves), a reducing
valve is normally open and blocks flow at set pressure.

Figure 14-13. Pilot-operated reducing valve

The normally closed direct-acting relief valve in the pilot section traps fluid from the
reduced-pressure outlet port through the control orifice on top of the spool when
pressure is below its setting. The spool stays in the normally open position because
pressure on both ends balances it hydraulically while the light spring keeps it pushed
down. As pressure at the reduced-pressure outlet port continues to increase, it finally
starts to open the direct-acting relief valve in the pilot section. Some fluid then flows to
tank through the drain port. When flow through the direct-acting relief valve is more
that the control orifice can handle, pressure on top of the spool drops and pressure on
the bottom of the spool pushes it closed. The spool never closes completely because
there is flow through the drain port anytime pressure at the outlet is lower than at the
inlet. Drain port flow amounts to about 60 to 90 in.3/min. This flow is all wasted energy
and it can cause a system to overheat if more reducing valves are installed than
necessary. When pressure drops below the direct-acting relief valve’s setting in the pilot
section, the valve closes and forces the spool to the open position.

A reducing valve is normally open so it appears reverse flow should not be a problem.
However, when the valve is working, it is almost closed — and it can be held closed by
back flow when the actuator starts to return. Anytime a reducing valve must pass
reverse flow, select a valve with an integral bypass check valve to eliminate the
possibility of blocked return flow.

It also is very important to have a free-flow drain port with very low (or even no)
backpressure. Backpressure in the drain port adds to the setting of the direct-acting
relief valve and can cause erratic results when drain pressure fluctuates. (Our next e-
book, Fluid Power Circuits Explained, discusses how this drain port can be used
advantageously in a dual-pressure circuit. This book will be launched in the next few
months.)

Figure 14-14. Dual-pressure hydraulic circuit using a relief valve

The modified circuit in Fig, 14-14 allows two pressures without lowering system
pressure (as happened in Figure 14-13). A pressure-reducing valve in place of Relief
Valve 2 makes it possible to set pressure for Cylinder 2 without affecting pressure
at Cylinder 1. This reducing valve never has reverse flow so a bypass check valve is not
required.

When it is working, a reducing valve is nearly closed and will pass very little reverse
drain flow unless it has a bypass check valve. Even then, reverse flow must be at a
pressure greater than that at the inlet. If this much pressure cannot be tolerated, use the
reducing/relieving valve depicted in Figure 14-15.
Figure 14-15. Pilot-operated reducing/relieving valves

Reducing/relieving valves function exactly like reducing valves — until an external force
starts to increase pressure at the reduced-pressure outlet above the pressure set by the
pilot section. When outlet pressure is 4 to 6% above set pressure, the spool moves up
until the outlet is connected to tank. Any fluid at pressure above set pressure returns to
tank, so outlet pressure does not continue to climb. Tank flow comes only from the
reduced-pressure outlet, not from the pump through the inlet. When excess pressure at
the outlet drops, the reducing/relieving valve continues to perform its reducing
function.

Note that the left cutaway view has an internal drain for the pilot section. This saves
connecting a separate drain line for pilot flow. However, when backpressure in the tank
line is high or may fluctuate due to other return functions, it adds to the pilot-section
setting and can elevate pressure at the reduced-pressure outlet above allowable rates.
When tank-line backpressure may be high or when pressure fluctuations cannot be
tolerated, use a valve with an external drain. When reverse flow is necessary, specify a
model with an integral bypass check valve for piping convenience.

Figure 14-16. Modular sequence, counterbalance, and reducing/relieving valves

Figure 14-16 shows most of the modular valve configurations for sequence,
counterbalance, and reducing valves. Modular valves simplify piping and eliminate
many connections that can generate backpressure or add potential leakage points.
RELATED
CHAPTER 13: Flow Controls and Flow Dividers

QUIZ on Chapter 13 Table of Contents Answers to Quiz 13

Bud Trinkel | Mar 24, 2007

It is necessary in many applications to vary the speed of an actuator. One method of


controlling an actuator’s speed is by using a variable-displacement pump. This works
well for a circuit with a single actuator or in multi-actuator circuits where only one
actuator moves at a time. However, most circuits that need actuator-speed control have
multiple actuators, and some of them operate simultaneously. For most circuits, a
variable orifice, called a needle valve or flow control. Fixed orifices may be used in
some cases.

Flow-control valves

A non-compensated flow control passes more or less fluid as pressure increases and
decreases. This is because more fluid can pass through a certain size orifice when
pressure drop across the orifice increases. Figure 13-1 shows non-compensated flow
devices in symbol and cutaway form. At the top are non-compensated fixed-orifice
inline flow controls for tamper-proof applications. These can be purchased as in-line
valves or they could be a drilled plug or insert located in a pipe fitting or valve port.

Flow through standard orifices is affected by viscosity changes in the fluid, while flow
through knife-edge (or sharp-edge) orifices changes very little when fluid viscosity
changes from thin to thick. A knife-edge orifice is the style used on most valves that are
designated as temperature compensated. (A classic example of a non-compensated fixed
orifice with a bypass check is the orificed check valve shown in Figure 10-2.)
Fig. 13-2. Pressure- and temperature-compensated
flow control

Pressure-compensated flow control valves are used with actuators that must
move at a constant rate, regardless of pressure. A pressure-compensated flow control
cutaway view and symbols depicted in Figure 13-2. The needle valve section of a
pressure-compensated flow control is the same as any flow control. The difference is the
addition of a compensator spool that can move to restrict Inlet flow at the compensating
orifice. The compensator spool is held open by a 100- to 150-psi bias spring that sets
pressure drop across the knife-edge orifice.

Flow from the inlet goes through the compensating orifice, past the compensator spool,
and out through the knife-edge orifice. A drilled passage ports Inlet fluid to the right end
of the compensator spool, which forces the spool to the left when pressure tries to go
above 100 to 150 psi at gauge PG01. After pressure reaches or goes above 100 to 150 psi,
the compensator spool moves to the left and restricts flow to the knife-edge orifice flow
control. Pressure at gauge PG01 never goes above 100 to 150 psi (plus any backpressure
at the outlet). Pressure at the outlet is ported to the bias-spring chamber and increases
the spring force. The compensator spool assures that pressure drop across the knife-
edge orifice flow control stays at a constant 100 to 150 psi. With a constant pressure
drop, flow stays the same regardless of inlet or outlet fluctuations.
Pressure-compensated flow controls are four to eight times more expensive than
standard controls so they should only be applied to actuators that must move
consistently.

The no-jump option is an adjusting screw that holds the compensator spool within a few
tenths of an inch of its operating position. This is an especially important option when
the valve is oversize for the present flow setting. A compensator spool without a stroke
limiter may close and open violently until it stabilizes and sets pressure drop for the
orifice. During this time the actuator also moves erratically.

The two symbols represent the American National Standards Institute (ANSI) and the
International Standards Organization (ISO) way of indicating that the valve is pressure
compensated. The arrow indicating pressure compensation is easier to distinguish in the
ANSI symbol — especially when the schematic drawing has been reduced to fit into a
machine's documentation book.

Fig. 13-3. Three-port flow control

Three-port flow control valves are mainly used in fixed-volume pump circuits to
save energy. (See the load-sensing pump circuit explained in Chapter 8.) If 20 gpm of
fluid enters the Inlet and the flow control is set at 12 gpm, 8 gpm goes to tank as wasted
energy. With a conventional relief valve setup, pressure between the pump and flow
control would be maximum. With the 3-port flow control, pressure in this portion of the
circuit is whatever it takes to move the actuator plus bias-spring force. (Bias-spring
force is usually 70 to 125 lb.) An outlet pressure of 200 psi gives a pressure of 270 psi
between the pump and the flow control. All fluid going to tank is discharged at 270 psi,
not 2000 psi. This takes place because the sensing line sends feedback to the pressure-
control side of the relief valve, allowing it to open at load pressure plus bias-spring force.
Pressure between the pump and flow control constantly changes with load variations.
When the load requires more than the maximum-pressure adjustment setting, the relief
valve opens and sends all pump flow to tank at maximum pressure.

A 3-port flow control is only effective with one actuator — or one actuator at a time. It
would not be useful on a pressure-compensated pump circuit because a load-sensing
circuit for this type pump would save even more energy. (See Chapter 8 for a load-
sensing circuit with a pressure-compensated pump.)

Fig. 13-4. Proportional flow control valve without


feedbackProportional flow-control valves are shown as cutaways and symbols for
proportional flow control valves that can electronically remotely control flow through a PLC or
other control. Many different designs of valves and controllers can control pneumatic or
hydraulic fluid. The design in Figure 13-4 uses a modified 2-way pilot-to-close poppet with a
drilled pilot passage to send inlet fluid behind it. A light spring holds the poppet closed when no
pressurized fluid is at the Inlet.

The armature controls a small normally closed poppet and shifts the signaled amount to
let fluid behind the pilot-to-close poppet leave faster than the pilot passage can supply
it. This causes a pressure imbalance that lets the pilot-to-close poppet open enough to
give the correct fluid flow. The flow rate is infinitely variable and can be controlled from
a variety of inputs.

Fig. 13-5. Proportional flow control valve with feedback

The cutaway and symbol in Figure 13-4 represents a valve that opens from a given
signal but may not always repeat a set flow from the same input. The feedback LVDT
added to the valve in Figure 13-5 assures that the pilot-to-close poppet always shifts
the same amount so it has the same size flow opening. However, pressure or viscosity
changes still affect actual flow, so a hydrostat is necessary when exact flow repeatability
is required. Many manufacturers make valves with a built-in hydrostat for pressure
compensation.

Meter-in flow control circuits

Figure 13-6 provides a schematic drawing of a meter-In flow control circuit restricting
fluid as it enters an actuator port. Meter-in circuits work well with hydraulic fluids, but
can give erratic action with air. Note that the cylinder is horizontally mounted, which
makes it a resistive load. Meter-in flow controls only work on resistive loads because a
running-away load can move the actuator faster than the circuit can fill it with fluid.
Fig. 13-6. Meter-in
flow control circuit

The left-hand circuit in Figure 13-6 is shown at rest with the pump running. Notice that
the check valves in the flow controls force fluid through the orifices as it enters the
cylinder and lets fluid bypass them as it leaves.

The right-hand circuit depicts conditions as the cylinder extends. The directional control
valve shifts to straight arrows and pump flow passes through the left-hand flow control
to the cylinder cap end at a controlled rate. Fluid leaving the cylinder rod end flows to
tank without restriction. The cylinder extends at a reduced speed (in a hydraulic circuit)
until it meets a resistance it can’t overcome or it bottoms out. With the non-
compensated valve shown, speed can vary as pressure fluctuates or viscosity changes.

While the cylinder is in motion, pressure at PG1 reads the setting of the relief valve or
pump compensator. The pressure at PG2 reads whatever it takes to move the load at any
point in the cycle. Pressures at PG3 and PG4 only read tank-line backpressure as the
cylinder extends.

It is obvious that if the cylinder had an external force pulling on it, it would extend
rapidly. Because fluid enters the cap end at a reduced flow rate, a vacuum void would
form there until the pump had time to fill it.

Meter-in flow controls can have a problem in pneumatic circuits. When fluid is directed
to the cylinder cap end, pressure at PG1 immediately rises to the regulator setting.
However, pressure at PG2 starts at zero and increases slowly. Until pressure at PG2 rises
enough to generate breakaway force, the cylinder does not move. At breakaway
pressure, the cylinder extends quickly and expanding air may cause it to lunge. Often,
the lunge forward moves the piston ahead of the incoming air and pressure drops back
below the breakaway level so the piston stops. Pressure starts to build again and the
lunge/stop scenario continues to the end of stroke. The meter-out circuit discussed next
is always the best choice to control air cylinders.

The circuits in Figure 13-7 show applications where a meter-in circuit is the only
choice for both pneumatics and hydraulics. On the left in Figure 13-7, a single-acting
pneumatic cylinder is mounted with the rod vertically up. The only way to control
extension speed is via a meter-in flow control. When retraction speed must be controlled
as well, a meter-out flow control also is necessary.
Fig. 13-7. Circuits
where meter-in flow control is required

The cylinder pictured on the right in Figure 13-7 is extending to perform an operation
prior to retracting or starting the cycle of another actuator. A signal to continue the cycle
can come from a pressure switch or a sequence valve. Either of these devices can be set
to give an output at any pressure. Usually they are set 50 to 150 psi below system
operating pressure for hydraulics, or 5 to 15 psi lower for air. The reason for meter-in
flow control is that pressure between the flow control and the cylinder normally stays
low until the cylinder contacts the workpiece. At work contact, the resulting pressure
buildup switches these pressure-actuated devices and starts the next sequence. Always
remember: a pressure switch or sequence valve does not directly indicate that the
actuator has reached a physical position. They only indicate that pressure has reached a
predetermined setting . . . not why it has.

Other circuits that require meter-in flow controls are the load-sensing pump circuits in
Chapter 8.

Meter-out flow control circuits

Figure 13-8 shows a schematic drawing of a meter-out flow control circuit that
restricts fluid as it leaves an actuator port. Meter-out circuits work well with both
hydraulic and pneumatic actuators. Cylinder-mounting attitude is not important
because outlet flow is restricted and an actuator cannot run away. Meter-out flow
controls work on resistive loads or running away loads because the actuator can never
move faster than the fluid leaving it allows.

Fig. 13-8. Meter-out


flow control circuit
The left-hand circuit in Figure 13-8 is shown at rest with the pump running. Notice
how the check valves in the flow controls allow fluid to bypass the orifices and freely
enter the cylinder. As fluid leaves the cylinder, it is forced through the orifices at a set
rate. The only gauge showing pressure is PG3 because the load on the cylinder rod is
inducing pressure at the valve’s blocked port.

The right-hand circuit shows conditions when the cylinder is extending. The directional
control valve shifts to straight arrows and pump flow bypasses the upper flow control to
go to the cylinder cap end. Fluid leaving the cylinder rod end is held back before it goes
to tank -- even with an external load trying to move it. The cylinder extends at a reduced
speed in both hydraulic and pneumatic circuits until it meets a resistance it can’t
overcome or it bottoms out. With the non-compensated valve shown, speed can vary as
pressure fluctuates or viscosity changes in a hydraulic system. (There are no pressure-
compensated flow controls for pneumatic circuits.)

While the cylinder is in motion, gauges PG1 and PG2 read the relief valve or pump
compensator setting. Gauge PG4 reads tank backpressure. Gauge PG3 reads load-
induced pressure plus the pressure from cap-area-to-rod-area intensification. This
intensified pressure could be 1.2 to 2 times the cap-end pressure, or higher, depending
on the rod size.

Meter-out flow controls work equally well in pneumatic circuits when the load is
constant. Changing loads can cause the actuator to stop and/or lunge under certain
circumstances. (For a more extensive coverage of flow control circuits and situations
that can arise with them, see our second e-book entitled "Fluid Power Circuits
Explained," which will be launched on hydraulicspneumatics.com in the coming
months.

Bleed-off flow control circuits

Bleed-off flow control circuits are found only in hydraulic systems and normally only in
those with fixed-volume pumps. There is little or no advantage to using this type flow
control with pressure-compensated pumps. Figure 13-9 shows a bleed-off circuit at
rest with the pump running. A needle valve’s inlet is teed into a line going to the cylinder
and its outlet is connected to tank. The circuit only works with one actuator moving at a
time because all pump flow goes to the presently operating function. Like a meter-in
circuit, it only works with resistive loads because it controls fluid into the actuator. The
main plus for this type speed control is it saves energy while using a fixed-volume pump
with low-pressure travel forces.

Fig. 13-9. Bleed-off


flow control circuit
When the directional valve in Figure 13-9 shifts, all pump flow passes through it and
toward the actuator. On the way to the actuator, part of the flow is bled off to tank, so
the actuator does not reach full speed. Pressure at PG1 only rises to whatever it takes to
move the actuator and its load, so excess flow goes to tank at low pressure. (When using
a fixed-volume pump and a meter-in or meter-out circuit, excess flow also goes to tank,
but at relief valve pressure.) Many circuits only perform work at the end of stroke so this
flow control system saves energy while the actuator moves to and from the work
position, yet still gives good speed control.

Some words of caution:

 Pressure in the actuator during traverse time must be higher than the pressure in the path to
tank, so fluid will flow to tank.
 Because pressure may change during traverse time (especially when the actuator contacts the
workpiece), use a pressure-compensated needle valve so flow to tank remains constant.
 Even with a pressure-compensated needle valve, actuator speed will be inconsistent. Pump
and/or actuator efficiency allows bypass that directly affects flow to the actuator not bleed-off to
tank.

Pressure-compensated flow control valve applications

When pressure drop across an orifice changes, flow through the orifice also changes. As
pressure drop increases, flow increases, and as pressure drop decreases, flow decreases.
Because of this fact, if pressure drop across an orifice were constant, regardless of
upstream and downstream pressure fluctuations, then flow through it would stay the
same. A pressure-compensated flow control valve (such as the one shown in Figure 13-
2) automatically maintains a constant pressure drop across the orifice. There is a short
discussion on pressure-compensated flow control valves on page 13-1, but a valve in
cutaway form is applied to a bleed-off circuit in Figure 13-10.

Fig. 13-10. Bleed-off


flow control circuit with positive-displacement pump and pressure-compensated flow control
valve
In the bleed-off circuit, fluid from the directional control valve is sent to the cylinder to
start it extending. Because the circuit has a fixed-volume pump and needs speed control,
a bleed-off flow control is used to save energy. Instead of controlling flow to or from the
actuator, excess flow is bled to tank across a pressure-compensated flow control at
whatever pressure it takes to move the fluid. A meter-in or meter-out flow control circuit
would send excess flow to tank across the relief valve at maximum pressure – wasting a
lot more energy.

The reason for using a pressure-compensated flow control is that pressure will fluctuate
as the actuator moves toward the work piece and the flow to tank from a non-
compensated flow control would change continuously. As a result, actuator speed could
vary considerably while it moves. With a pressure-compensated flow control, flow to
tank is constant, but actuator speed could still change due to pump efficiency as
pressure increases or decreases. Any speed change from pump efficiency is present but
practically imperceptible.

In the Figure 10-13 circuit, a 10-gpm pump sends 7 gpm to the cylinder and 3 gpm to
tank. Fluid entering the pressure-compensated flow control passes by the compensator
spool and flows on to the variable knife-edge orifice, which is set at 3 gpm. The variable
knife-edge orifice restricts flow and creates backpressure in the incoming fluid. When
backpressure reaches (and attempts to exceed) 125 psi, fluid in the inlet-pressure pilot
line forces the compensator spool to the right. This restricts flow at the compensating
orifice. After the compensator spool settles in at its 125-psi bias-spring setting, pressure
at PG3 reaches 125 psi and stays there. This means that pressure drop across the
variable knife-edge orifice is 125 psi. As the cylinder continues to move and pressure
at PG1 and PG2 increases or decreases, pressure at PG4 stays at 125 psi and flow is
constant. The cylinder moves at the same speed whether pressure is at or above 125 psi,
and as much as 125 psi below the maximum pressure setting.
Fig. 13-11. Meter-in
flow control circuit with pressure-compensated pump and pressure-compensated flow
control valve

Figure 13-11 shows a pressure-compensated flow control in a meter-in circuit. Fluid


from the valve enters the flow control and is restricted. Backpressure from restricted
flow goes through the inlet-pressure pilot line and shifts the compensator spool to the
right, restricting flow to the variable knife-edge orifice. Backpressure from cylinder
resistance acts on the right end of the compensator spool through the outlet-pressure
pilot line and adds to the 125-psi bias-spring force. This action and interaction always
keeps pressure 125 psi higher at PG5 than at PG2. A constant pressure drop across the
orifice maintains a constant flow to the cylinder.

Fig. 13-12. Meter-out


flow control circuit with pressure-compensated pump and pressure-compensated flow
control valve

Figure 13-12 shows a pressure-compensated flow control in a meter-out circuit. Fluid


from the cylinder rod end enters the pressure-compensated flow control and is
restricted at the variable knife-edge orifice. Backpressure through the inlet-pressure
pilot line shifts the compensator spool to the right and restricts flow to the variable
knife-edge orifice. Pressure at PG5 settles in at 125 psi and flow stays the same across
the variable knife-edge orifice. Any backpressure from tank flow adds to the 125-psi
bias-spring force and increases pressure at PG5 so it always stays 125 psi above PG4.

Pressure-compensated flow control valves are as much as five times more expensive
than non-compensated models, so they should not be specified when accurate flow
control is not required.

Changes in fluid viscosity also cause flow fluctuations. Thick fluid flows more slowly
than thin fluid. A flow control valve without temperature compensation allows varying
flow from cool oil at startup to oil running at normal or high temperature. The most
common fix for viscosity variations is to use a knife-edge orifice. Knife-edge orifices
have no flats to slow fluid flow, so they produce little change in flow between thick and
thin fluids. Other devices to obtain constant flow with viscosity variations are available,
but they can be complex and may cause malfunctions.

A flow control in a hydraulic circuit always generates heat. Some pump and flow control
combinations produce a lot more heat and should be avoided if possible. The following
examples show different pump and flow control combinations and suggest how much
heat can be expected.
Fig. 13-13. Heat
generation in fixed-volume pump circuits with meter-in and meter-out flow controls

The fixed-volume pump and meter-in or meter-out flow control combination in Figure
13-13 is the worst-case situation. The example shows a cylinder stroking to the
workpiece with flow controls set at 3 gpm. A 10-gpm pump driven by a 5-hp electric
motor powers the circuit. Because it only takes 100 psi to move the cylinder while
traversing, a lot of heat-generating energy is wasted. This example is somewhat
exaggerated, but is not at all unheard of. Note the example only shows energy wasted on
the extension stroke. With a reduced-speed retraction stroke, heat generation could
almost double the figures shown.

The main generator of heat is the excess pump flow going across the relief valve at 1000
psi. The two circuits in Figure 13-14 show how to eliminate such wasted energy with a
different flow control circuit or a different pump. While the energy wasted across the
flow control valve is much less at these low flows, it still adds heat to a system. Also, the
amount of pressure drop may be lower than indicated here because some actuators
require more pressure to move them to and from the workpiece. Energy loss across a
flow control cannot be eliminated. The amount of loss depends on pressure drop and
flow rate across the orifice.

Fig. 13-14. Two flow


control circuits that reduce heat generation
The circuits in Figure 13-14 show a fixed-volume pump with a bleed-off circuit and a
pressure-compensated pump with a meter-in circuit. Both of these combinations save a
lot of energy (although not as much as the load-sensing circuit that was shown
in Figure 8-27). This type of flow control circuit wastes the least energy possible when
using flow controls for speed control.

Fluid flow dividers

The flow divider in Figure 13-15 is called a priority flow divider because it splits pump
flow into a fixed controlled-flow (CF) outlet and sends excess fluid out an excess flow
(EF) port. Volume orifices (drilled as specified by the purchaser) preset fluid flow out of
the CF port. EF flow is any flow the pump produces over and above the controlled flow.
This type flow divider is often used on vehicle power steering, where an engine-driven
pump’s output may vary as rpm changes or as its flow is used for other functions. A
priority flow divider assures that the power steering always has ample fluid at any
engine speed or when other functions are active.
Fig. 13-15. Priority flow divider with relief
valve in priority leg

As fluid enters the valve, the path of least resistance leads through the controlled-flow-
volume orifices and out port CF. If pump flow is more than the volume orifices can pass,
pressure builds on the right end of the flow-control spool through the excess-flow pilot
line. When pressure rises enough to overcome the bias spring and any backpressure
from the steering circuit, the flow-control spool moves to the left, just enough to let
excess flow exit through port EF. Excess flow changes as pump flow varies, but flow to
port CF takes priority. A relief valve in port CF can be set for any pressure and has no
affect on pressure at port EF. The controlled-flow relief valve is required even when
maximum pressure is the same for both outlets.

Notice that controlled flow is pressure compensated. As pressure builds at port CF, it
pushes back against the excess-flow pilot-pressure pilot to maintain a constant pressure
drop across the volume orifices.

Priority flow dividers are also manufactured with adjustable flow for the priority port
and without a relief valve for circuits that already have one. (The symbol shown is
borrowed from a manufacturer's catalog because there is no standard symbol in ANSI or
ISO literature.)
Fig. 13-16. Spool-type flow divider for 50-50 split

The flow divider in Figure 13-16 is a spool-type divider that splits flow at any
predetermined rate according to the sizes of the drilled orifices. It is usually set up with
identical orifice sizes for a 50-50 split. This particular design does not allow reverse
flow, so bypass check valves are required when flow must return the same way it
entered.

Fluid entering the Inlet port goes left and right through orifices, then out outlets 1 and 2.
When either outlet encounters more backpressure than the other does, the high-
pressure side forces the spool towards the low-pressure side until pressures on both
sides equalize. Equal pressure drop across both orifices produces equal flow. (Most
manufacturers specify flow equality at ±5%.) Pressure differences at the two outlets
should be low because Inlet pressure always equals the highest outlet pressure -- which
means pressure drop across the low-pressure outlet wastes energy.

Spool-type flow dividers only split flow. When more than two outlets are required,
dividers must be used in series. A 50-50 split divider flowing into two more 50-50
dividers gives four equal outlets. A 66-33 divider into a 50-50 divider gives three equal
outlets. The flow divider/combiner in Figure 13-17 equalizes flow in both directions. It
can be used with double-acting actuators to synchronize speed in both directions of
travel. The spool in this divider is made in two sections with a connecting link that
allows the sections to move together in the closed condition (as shown) for combining,
or be spread by Inlet pressure when they are dividing. Springs at both ends of the spool
keep the sections together when pressure equalizes or is not present. Inlet orifices set
nominal flow, while outlet orifices control flow to or from an actuator.

Fig. 13-17. Spool-type flow divider/combiner with 50-50


split

Flow to the inlet-return port goes through the inlet orifices to split into two equal parts.
Pressure drop across the orifices causes the split spool to separate so the outlet orifices
are working at the outer edge of the outlet-return ports. When unequal pressures on its
ends shift the spool, flow is retarded to the low-pressure outlet port to keep it from
receiving too much fluid. When the actuator reverses, flow into the outlet-return ports
goes through the outlet orifices and on through the inlet orifices, causing the spool
sections to come together. Now the outlet orifices control return flow on the inner edge
of the outlet-return ports. They will retard flow from any actuator port that is trying to
run ahead.

Rotary flow dividers

A rotary (motor-type) flow divider is constructed from two or more hydraulic motors —
in a common housing — with a common shaft running through one set of gears on all
motor sets. There is a common Inlet to all motors and separate outlets. The motors are
usually gear-on-gear or gerotor design. Flow split is commonly 50-50 but many outlet
flow combinations are possible by changing gear or gerotor widths.
The cutaway view and symbol in Figure 13-18 pictures a 2-outlet 50-50 split gear-
motor-type flow divider. (There is no ISO or ANSI symbol for a motor flow divider so
the one shown in the figure is from a supplier’s catalog.) One gear from each motor set is
keyed to the common shaft, so both motors must turn at the same rate. If one motor
stalls, they both stop because of the common-shaft arrangement. Due to internal
clearances in the motor elements, there is some bypass flow that does not turn the
motors. As a result, the outlet flows are not always exactly equal . . . especially at high
outlet-pressure differences.

Fig. 13-18. Motor-type flow divider with 50-50 split

From Figure 13-18, it should be obvious that this flow divider does not have a priority
side like a spool-type flow divider does. Thus, when Inlet flow changes, it is always split
equally. The main advantage of motor-type over spool-type flow dividers is there is less
wasted energy when the outlets are not at or near the same pressure. If pressure at the
right outlet was 1500 psi and pressure at the left outlet was 300 psi, pressure at the inlet
would be 900 psi. Pressure at the inlet is always the average of the sum of the outlets.

This feature can be an asset or a problem. If one outlet meets resistance while the other
is flowing to tank, an inlet pressure of 2000 psi can result in the pressurized outlet
intensifying to 4000 psi. If pressure that high cannot be tolerated, a relief valve must be
installed at the outlets. On the other hand, intensification can allow a 1000-psi system
to produce 2000 psi to perform work -- similar to a hi-lo pump circuit. Note that while
pressure doubles, flow is halved through the high-pressure outlet.
Fig 13-19. Synchronizing circuit for 50-50 flow divider

Looking at Figure 13-18, it appears the motor flow divider is also a combiner. This is
partially true. The circuit in Figure 13-19 shows a motor flow divider synchronizing
two hydraulic motors. As the motors turn in right-hand rotation, they stay almost
perfectly synchronized. Pressure to each motor may vary but flow from each flow-
divider outlet remains near constant. If the directional control valve shifts to turn the
motors in left-hand rotation, the flow divider may get equal flow and the hydraulic
motors may stay synchronized. However, if one hydraulic motor meets more resistance
than it can overcome and stalls, all pump flow goes to the running hydraulic motor. The
second motor then turns twice as fast. During this scenario, one flow-divider motor
overspeeds while the opposite one cavitates. The only way to make sure both hydraulic
motors stay synchronized in both directions of rotation is to install motor flow dividers
at both valve ports.

Spool and motor flow dividers work reasonably well to synchronize circuits with
hydraulic motors and cylinders. However, because both devices do not divide flow
perfectly, the actuators they control will not stay perfectly synchronized. A high-
pressure difference at the divider's outlets is the worst problem; it can allow a 5 to 10%
lag in actuator position. This means that synchronizing circuits using flow dividers often
require some type of re-synchronizing valving to realign the actuators more exactly
when they stop at home position. (Due to internal bypass, actuators with short cycles
may re-synchronize themselves because the error is small.)

Fig. 13-20. Motor-type flow-divider circuit


with 50-50 split

Another design consideration is the intensification of pressure at the outlets of a motor


flow divider. The circuit in Figure 13-20 has two cylinders that are synchronized by a
motor flow divider. Because this circuit operates at 2000 psi, it is possible that pressure
at one cylinder could reach as much as 4000 psi due to intensification. Intensification
occurs when one cylinder is lightly loaded or has no load and the other one is loaded
heavily. In Figure 13-19, the load is shifted to one side of the platen -- making the
right-hand cylinder do all the work. Inlet pressure is at 2000 psi and the cylinders are
stalled. Pressure at the lightly loaded left-hand cylinder is 250 psi, so pressure at the
right-hand cylinder is 3750 psi. The intensification is due to energy transfer through the
motors in the flow divider. Because inlet pressure for both motors is 2000 psi, the
unused 1750 psi from the left side is transmitted through the common shaft and drives
the opposite motor to 3750 psi. (For other flow-divider circuits. see the author’s book,
“Fluid Power Circuits Explained,” available through the same outlet for this manual.)
Fig. 13-21. Symbols for modular flow
controls and flow dividers

Most flow control functions are available as modular or sandwich valves that mount
between directional control valves and a subplate. Figure 13-21 shows most of the
common configurations presently offered by fluid power suppliers. Although the
symbols show non-compensated flow controls, most configurations also are available
with pressure-compensated flow controls. Where a needle valve is shown, a flow control
with bypass may actually be installed. This is not a problem because there is never a
reason for flow reversal. Figure 13-21 also shows two modular flow dividers that are
available from one supplier. These modules are usually available in all valve sizes up to
D08 (¾-in. ports).
RELATED

SPONSORED CONTENT
CHAPTER 14: Sequence Valves and Reducing Valves

APR 08, 2007

CHAPTER 12: Infinitely Variable Directional Valves

Quiz on Chapter 12 Table of Contents Answers to Quiz 12

Bud Trinkel | Mar 03, 2007

Proportional and servovalves


Infinitely variable directional control valves

The directional control valves discussed so far in this series have all been configured to
either pass full flow or completely block flow. The only way to decrease flow through
these valves is by adding flow controls or by mechanically limiting movement of an
internal part.

The first infinitely variable valve available was the servovalve. Internal flow-modifying
parts could be moved to any position at any rate, so output from any port could be
varied at will. (Some call these valves infinitely variable 4-way flow controls.) The main
problem with servovalves was (and still is) that they require very clean fluid to keep
them operating effectively. Fluid from a standard well-maintained hydraulic circuit
contains enough contamination to cause most servovalves to fail in a matter of minutes
or only last a few hours at best. This meant that the original servovalves were tried and
removed from many machines that needed precise control but not at the perceived cost
of cleaning up the hydraulic oil.

Why use infinitely variable valves?


Some actuators must move at a precise speed, stop at a close-tolerance position, or
produce a very accurate force to perform the work for which they were designed. With
the proper input signals and feedback devices, proportional or servovalves can make an
actuator perform any or all these functions flawlessly.

Rolling mills turn out sheet consistently to a tolerance of ±0.0005 in. at sheet speeds of
2000 to 5000 feet per minute. Hydraulic cylinders controlled by servovalves maintain
the proper force and position the rolls precisely from feedback signals sent by sensors
that measure metal thickness, cylinder force, and position. Airline pilots train in
simulators moved by hydraulic cylinders so precisely that the pilots get the feel of
landing gear raising and locking in position. Even entertainment rides use servovalves
to make passengers think they are in 20-ft waves when they are actually in an enclosed
articulated room in a shopping mall.

For less precise movement, there are proportional valves that mimic the output of
servovalves but respond more slowly. They are less expensive than servovalves and more
contamination tolerant, so they have replaced cam valves and other mechanical devices
used to get smooth motions.

Hydraulic proportional directional control valves

The symbol and cutaway view in Figure 12-1represent a direct-acting proportional


valve that handles flows as high as 10 to 30 gpm in D03- or D05-size valve interfaces.
Proportional valves use the same interface standards as NFPA and ISO directional
valves so they can be installed in a circuit without having to change the piping.
Fig. 12-1. Direct-acting proportional valve

Physically, proportional valves appear the same as their on/off solenoid counterparts.
The big difference is in the way their solenoid coils perform. Proportional coils operate
on DC current and produce varying force with varying voltage. The symbol shows the
solenoid slash in the operator box with a sloping arrow through the slash. This indicates
the solenoid has variable force that moves the spool more or less as voltage increases
and falls. The other indication on the symbol that shows the spool is infinitely variable is
the parallel lines down both sides of the boxes. Proportional valves operate similarly to
manual valves, but they use electronics instead of hand power.

To eliminate flow lag from spool overlap, most manufacturers cut vee notches or use
some similar method that allows some flow to pass as soon as the spool moves. Vee
notches also give smooth flow buildup until the spool moves through the land overlap.

Proportional valves only have two center configurations (as shown by the symbols
in Figure 12-1). This means that pressure-compensated pumps with accumulators
normally power circuits with proportional valves. The circuits are pressurized at all
times to produce fast response from an actuator when motion is called for. (A pressure-
compensated circuit also wastes the least amount of energy when throttling flow.)

The valve in Figure 12-1 depends on a certain voltage to move the spool a certain
distance to pass a certain flow. This works reasonably well, but is not accurate over a
broad range of pressures, flows, and temperatures. Most valves of this design are used to
smoothly accelerate and/or decelerate an actuator. The spool is electronically controlled
to shift over a period of time to increase flow at a controlled rate. Spool-shift speed can
be controlled electronically as it opens and closes to give smooth acceleration and
deceleration. Spool-shift distance can also be limited electronically to set a maximum
speed when required.

Fig. 12-2. Direct-acting proportional


directional control valve with spool-position feedback transducer

To give better spool control, a linear variable-displacement transducer (LVDT) is


added to the basic valve. The cutaway view and symbol in Figure 12-2 represent a
direct-acting valve with an LVDT. Feedback from the LVDT tells the electronic
controller the spool’s position and makes sure it goes to the same place when it receives
the same signal. With this arrangement, the spool always shifts to an exact location and
opens the same size orifice so it can pass the same flow when pressure drop and
viscosity stay the same. Control of flow is more accurate with an LVDT but pressure
drop does not stay constant and viscosity often changes throughout the day so speed
variations are still apparent. Such flow variations caused by system pressure fluctuation
can almost be eliminated by the addition of a hydrostat module in port P as shown
in Figure 12-3.
Fig. 12-3. Direct-acting proportional
directional control valve with LVDT and pressure-compensating hydrostat module

A hydrostat is simply a pressure-reducing valve set to hold downstream pressure in a


100- to 150-psi range. However, a hydrostat has a pilot line from a shuttle valve that
reads downstream pressure at ports A and B, then feeds it back to the bias-spring end of
the spool that controls the 100 to 150 psi. The function of a hydrostat is to maintain a
constant pressure drop across the spool orifice so flow stays constant regardless of
changes in system pressure. (For a thorough understanding of how pressure
compensation works, see the pressure-compensated flow control valve section in
Chapter 13.)

With the addition of a hydrostat, actuator speed is controlled as accurately as possible


without a closed-loop electronic circuit that reads speed and modifies spool position.
Closed-loop electronic circuits are used with proportional valve systems, but they only
give nominal control. When accurate control is required, use servovalves with closed-
loop electronics.

The symbol and cutaway view in Figure 12-4 is for a proportional valve that only
controls flow. Such valves are commonly called throttle valves because they are not
pressure compensated unless a hydrostat module is added.
Fig. 12-4. Direct-acting proportional throttle
valve with spool-position feedback transducer

Basic operation is identical to the proportional control valves just discussed. The only
difference is they have a single solenoid and may be piped with dual flow (as shown).
Dual-flow piping allows a given size valve to pass twice the volume at the same pressure
drop. It can be used with any 4-way directional control valve with one precaution: the
valve must be capable of handling maximum system pressure in its tank port. Many wet-
armature valves will not operate at full rated pressure at their tank port. Check the
supplier’s catalog to see what maximum tank line pressure is allowed. Air-gap solenoid
valves and solenoid pilot-operated valves with external drains normally allow full rated
pressure in the tank port.

The circuit in Figure 12-5 shows a possible use for a proportional throttle valve. The
vertical-down acting cylinder with a platen needs speed, acceleration, and deceleration
control. This could be done with a 4-way proportional valve, but the circuit uses an
inline or screw-in cartridge valve that is not directly replaceable. Adding a proportional
throttle valve to the tank line of the present 4-way circuit can give the required control
without extensive piping changes. The circuit is shown using a single flow path for low
volume. A dual flow path setup (like the one in Figure 12-4) would allow as much as
twice the flow. As stated earlier, make sure the 4-way directional control valve can
accept tank-line backpressure without damaging it or causing a malfunction.
Fig. 12-5. Typical circuit using a direct-acting
proportional throttle valve with spool-position feedback transducers

If the cylinder had a resistive load, the proportional throttle valve could be placed in the
pump line of the 4-way as a meter-in flow-control circuit. A meter-in circuit would not
damage the directional control valve or cause it to malfunction. The circuit in Figure
12-5could have a counterbalance valve in the rod-end line to make it resistive. (See
Chapter 14 for counterbalance valve operation and applications.)

The platen would start to move at a controlled rate when the 4-way valve shifts and the
proportional throttle valve is signaled to open slowly. The shift time for the throttle
valve determines the acceleration time, while shift travel distance determines maximum
cylinder speed. Cylinder speed would be infinitely variable to match any production
need.

Near the end of the stroke, a slowdown limit switch would signal the proportional
throttle valve to start shifting back to its closed position. The proportional throttle valve
would close at a controlled rate and flow from the cylinder would be retarded smoothly.
When the cylinder slows sufficiently, it contacts the end-of-stroke limit switch and the
4-way directional control valve shifts to center to stop it.
The circuit in Figure 12-5 would only hold position if some retract signal was applied
when stopped. This is due to internal leakage by the spool of the 4-way directional
control valve and the proportional throttle valve. Another option would be to use a float-
center directional control valve and a counterbalance valve.

Proportional valves for flows higher than 25-to-30 gpm use solenoid pilot-operation
similar to conventional directional control valves. A small pilot-operated valve receives a
signal and then sends hydraulic oil to proportionally move a larger control spool that
controls actuator movement.

Fig. 12-6. Solenoid pilot-operated


proportional directional control valve with spool-position feedback transducers

The cutaway view and symbol in Figure 12-6depict a typical solenoid-pilot valve
arrangement. A reducing valve module, between the pilot operator and the pilot-
operated valve, keeps maximum pilot pressure below 200 psi. A proportional Input to
one of the coils on the pilot operator directs flow to the spool of the pilot-operated valve
and shifts it against a spring. As pressure against the spool increases, it shifts farther
and sends more flow to the actuator. Feedback signals from both spools tell the
electronic controls that the command has been carried out. A vee-notched spool allows
flow to increase at a smooth rate so actuator speed is consistent throughout the speed
range.

Ports for internal or external pilot X or drain Y provide options for these control lines to
meet a particular requirement.

The complete symbol is shown in Figure 12-6. (For the simplified symbol, see Chapter
4.)

Flows up to 200 gpm are common for a D10-size proportional valve. For higher flows,
use slip-in cartridge valves (discussed in Chapter 11) with proportional operators.

Hydraulic servo directional control valves

Proportional control valves are infinitely variable but they are neither highly responsive
nor capable of handling minute flow changes rapidly and accurately. On the other hand,
servovalves easily meet both of these requirements . . . but at a cost. They are more
expensive than proportional valves, they require super-clean fluid, and they need extra
electronics to exploit their full capabilities.

The three common servovalve types are flapper, jet pipe, and mechanical. Each design
has advantages as far as operation accuracy, leakage, contamination tolerance, and
price. They range in flow capacity from less than 1 gpm to more than 1000 gpm. Most
manufacturers make valves that operate at 3000 psi, but some offer valves at 5000 psi.

The main difference between proportional and servovalve circuit design is that servo
systems have a method of feedback that assures that the actuator is doing what the
controller tells it to do. A super-simple form of servo control would be a backhoe
operator moving manual valves to cause a bucket to move toward him at a given rate.
Feedback from the operator’s eyes would tell his hands when and how far to move the
levers to give more or less flow to maintain the action he wants. Other familiar
mechanical feedback examples are hydraulic driven power steering and hydraulic power
brakes on a vehicle.

The circuit in Figure 12-7 is an example of a working circuit with mechanical feedback
that controls a hydraulic press. The operator needs to have a feel for the motion of a
platen as it cuts through some tubing. Originally this was done with an arbor press, but
it was hard for the operator to keep up with production due to the physical exertion.
Now all the operator has to do physically is overcome the spring force of the manually
controlled mobile directional valve to make the platen move. As the platen moves, the
directional valve body also moves, so the operator has to keep moving the lever to
advance or retract the platen. Notice the mechanical link between the valve body and the
cylinder rod that moves the valve body at the same rate the cylinder rod moves. The
operator now has a hydraulic force multiplier that gives some feel to what he is doing.

Fig. 12-7. Simple mechanical servo system


for force multiplication
The reason for using a mobile-type valve is because those valves have less spool overlap
and the spool has notches cut in it. The notches pass a small flow almost immediately
when the spool moves. That flow increases in proportion to spool movement.

Most industrial applications use feedback from electronic linear, rotary, or force
transducers. A transducer is a device that produces an electrical signal in direct relation
to a position, force, or speed.

Linear potentiometers work for short strokes (12 in. or less). Longer strokes require a
device such as a Temposonics transducer. In either case, these devices feed a precise
position or speed indication back to an electronic controller.

For rotary motion, an encoder or similar device that produces multiple pulses per
revolution sends a signal about rpm or angle of rotation to the controller.

When information about force is required, a load cell sends the data to the controller.

With these very accurate feedback devices and a fast-response servovalve, an actuator’s
position, speed, and/or force can be repeatedly established within an extremely close
range. Electronics provides the accuracy while hydraulics provides the force via a super-
responsive servovalve.

Fig. 12-8. Rotary-drive servovalve


The cutaway view and symbol in Figure 12-8 show a less-responsive but more
contamination-tolerant servovalve. There are other mechanical ways of driving the
spool. The valve in Figure 12-9 uses a rotary drive and an eccentric to move the spool
left or right to an infinite number of positions. Because the drive is quite strong and
there are no orifices to clog, this valve can operate with fluid that meets ISO Code 4406
20/16/13.

Notice the difference in design between spools in proportional valves and servovalves.
Most proportional valves use spools with overlap and some sort of notches that pass
flow while moving out of overlap. A servovalve has no overlap or underlap of the spool
lands to the body lands. (One manufacturer calls it “Critical lap” because all points
blocking fluid cannot move without passing flow.) This spool design makes the valve
very responsive (as well as very expensive and prone to above average bypass).
Servovalve spools and bodies always come in matched sets because of their close fit and
four points of land-to-land match.

The rotary-drive eccentric valve pictured in Figure 12-8 has fast, controllable spool
movement from a rotary drive that incorporates a feedback loop. When the drive
receives a signal to move the spool to pass a certain flow, a position feedback output
sends a signal back to the controller when the motion is complete. There is still feedback
from the actuator that what was commanded is happening, so spool position can be
changed via the electronic feedback and controller as necessary.

Several factors determine when a given input will not produce the desired actuator
output. The main factor is actuator load. As load changes, input force must change -- by
allowing more or less fluid into the circuit. Fluid viscosity also has an effect, so the flow
path must be reduced as viscosity lowers and enlarged as viscosity rises. Then there is
system pressure. As pressure fluctuates, flow across the spool orifice changes. The
higher the pressure drop, the greater the flow. Because a servovalve circuit has feedback
from the actuator, it can adjust flow or pressure to match system changes continuously.
Figure 12-9 shows another mechanically driven servovalve. This setup works for
cylinders and hydraulic motors, but must be directly attached to the actuator (as shown)
or driven by it with a toothed belt and pulleys. This is a very contamination-tolerant
valve arrangement because the stepper motor is quite strong and there are no small
orifices to clog. The spool has no overlap or underlap so any movement immediately
initiates fluid flow to and from the cylinder. The piston and rod cannot rotate so the
feedback screw turns as the piston extends or retracts.

Fig. 12-9. Stepper-


motor-driven servovalve with mechanical feedback
As the stepper motor turns, a threaded rod inside the threaded spool moves the spool to
direct fluid to extend or retract the piston. When the cylinder piston moves, the
feedback screw turns the spool back on the threaded rod to counteract the stepper
motor shift. When the stepper motor turns, the piston moves at a speed proportional to
the stepper motor’s rpm. When the stepper motor stops, the piston catches up and stops
also. Manufacturers claim a tolerance of ±0.001 in. repeatability. If an external force
tries to push or pull the piston out of place, the feedback screw shifts the spool and fluid
starts resisting movement.

From the foregoing explanation, it is easy to see how this valve – when attached to a
hydraulic motor -- would give an exact number of turns and repeatedly cause the motor
to stop at exactly the same place. This will happens even if the hydraulic motor has
internal leakage. The only time the actuator gets out of place is when it cannot overcome
the load and stalls. The stepper motor continues to receive pulses, but it also stalls when
the spool shifts all the way. The stepper motor received a signal that should have placed
the actuator at a certain distance but the actuator did not get there because of
insufficient force. When the valve reverses, it starts its motion from the wrong point and
will overshoot home position as it returns. A limit switch at the home position can alert
the controller that there is a problem when the actuator overshoots -- and prevent the
machine from producing scrap.

The jet-pipe servovalve pictured in Figure 12-10 also tolerates contamination due to a
control orifice that is large enough to pass large particles. Pilot oil is tapped off the
system fluid inlet, sent through a coarse filter, and on to the jet pipe that terminates in
an orifice. The orifice outlet is centered over the inlet of two passages that terminate at
each end of a critical-lap spool. Flow into these passages puts equal pressure on both
ends of the spool as the feedback wire holds it centered. Current signals to the coils
cause the armature to rotate and shift more of the output of the jet pipe to one
passageway than the other. Pressure increases on one end of the spool and decreases on
the other end. As the spool shifts, it starts to pass flow to the actuator at a rate set by the
input electrical signal.

Fig. 12-10. Jet-pipe servovalve

When the jet pipe shifts to the left, the spool moves to the right. At the same time, the
feedback wire also moves to the right, pulling the jet pipe nozzle back to center and
stopping spool movement. A given input to the coils electromechanically shifts the
armature that moves the jet pipe. This moves the spool hydraulically and forces the jet
pipe back to center mechanically through the feedback wire.

A measured electrical input to a servovalve produces a fixed flow output, similar to a


proportional valve. This control alone does not give much better control than a
proportional circuit even though the valve is more responsive. There is still no
compensation for viscosity or pressure changes that can cause the actuator’s speed to
fluctuate. To overcome this problem, some sort of electronic feedback from the actuator
is necessary. The feedback signal through an electronic circuit board modifies the signal
to the servovalve to make the actuator perform as planned. Actually, the electronics do
the work as long as the valve can respond quickly enough to keep everything working at
the correct rate. This means the spool must be free enough to move easily without
excessive bypass. Anytime the spool moves, it should pass flow to and from the actuator.
The valve in Figure 12-10 is considered a 2-stage valve. The first stage is electronic and
receives an electronic input signal, while the second stage is fluid powered by a
hydraulic signal.

Fig. 12-11. Flapper-design servovalve

The jet-pipe servovalve depends on clean oil for long trouble-free operation -- not as
clean as the requirement for the flapper-valve design discussed next, but clean enough
to prevent the jet-pipe nozzle from clogging. It is obvious that once nozzle flow is
retarded enough or stops, the valve loses all ability to control flow to the actuator.

The most responsive and accurate servovalve design is the flapper valve, shown
in Figure 12-11. This design is the least tolerant of contamination because it depends
on very small orifices for fast response with minimal wasted energy. It is called a flapper
valve because the element that holds equal pressure on both ends of the spool at rest
reminds one of a flapping device. It is a 2-stage valve with an electronically controlled
torque motor as the first stage and a pilot-operated spool as the second stage. As in all
servovalves, the spool has no overlap or underlap that would make it sluggish or bypass
a lot of fluid unnecessarily.
Fluid from the pump inlet is tapped off through rather-coarse filter elements, passes
through orifices past both ends of the spool, goes on to nozzles, and out to the return
line. The orifice diameters are slightly larger than the nozzle diameters, so there is a
pressure buildup at both ends of the spool. A feedback wire attached to the flapper
terminates in a ball end that sits in a very close-fit slot in the spool. A sleeve around the
spool can be moved left or right by a null adjustment to align the spool and body lands
perfectly when the valve is first installed. (Usually null adjustment is only required at
startup of the valve.)

The null adjustment usually is a hexagonal wrench fitting attached to an eccentric pin
located in the sleeve slot. With the null adjustment centered, turning it one round moves
the sleeve from center to full right, back to center to full left, and back to center. If the
valve cannot be nulled within one rotation of the null adjustment, replace it and send it
in for repair. This usually indicates a clogged orifice or nozzle controlling one end of the
spool.

Unplug the electrical supply to the valve before setting null. Start the pump and watch
for actuator movement. If the actuator moves, loosen the null lock screw and carefully
turn it. Observe whether the actuator slows or picks up speed. A nulled valve stops
actuator movement because the forces on both sides are equal. High-flow 3-stage valves
cannot be nulled to the point of stopping an actuator due to the piloted spool slipping by
the stop-flow position as the pilot operator is adjusted. When null is set, lock the null
screw and reattach the electrical plug.

Turning the null screw with the electric plug detached is one way of moving an actuator
manually. This might be done to prove the valve is working properly and the problem is
electrical.

When the torque-motor coils receive a current signal, the armature rotates clockwise or
counter-clockwise and pushes the flapper closer to one nozzle and farther away from the
opposite one. This allows pressure to increase at one end of the spool and decrease at
the other. The spool then starts to move away from the higher pressure. If the armature
turns clockwise, pressure builds on the left end of the spool and it moves to the right, as
shown in the left cutaway view of Figure 12-12.

Fig. 12-12. Flapper-type


servovalve shifting from an electrical input signal

As the spool moves to the right, it also drives the feedback wire to the right. The
feedback wire is strong enough to overcome armature force and pull the flapper back to
center. After the flapper centers, pressure is equal on both ends of the spool and it stops.
More current to the coils causes more rotation and additional spool shift until the
feedback wire again centers the flapper.
From the foregoing explanation, it is obvious why this valve needs clean oil. If an orifice
or nozzle clogs, the spool shifts all the way to one end and the actuator moves until it
runs into a resistance it can’t overcome. Also, the spool must start shifting at a very low
pressure drop across it to keep response high. Contaminated fluid can cause sticking
and require high differential shifting pressure that makes spool movement erratic.

For flows above 60 to 80 gpm, a 3-stage servovalve is required. It consists of a small 2-


stage pilot operating a large pilot-operated spool, as depicted in Figure 12-13. The 2-
stage valve operates as just explained, but its output goes to move a pilot-operated spool
in the third stage to a precise position to control high flow to large actuators.

An LVDT signals the electronic control circuit that the pilot-operated spool is where it
was signaled to go. After receiving that position signal, the 2-stage valve shifts to no flow
or whatever flow it takes to keep the pilot-operated spool in place.

A 3-stage valve also depends on feedback signals from the actuator to modify the input
signal when the action is not in compliance with the command. This makes 3-stage
valves very accurate controllers of large cylinders.

Pneumatic proportional and servo directional control


valves

Since the late ‘80s, several companies have been controlling air cylinders with open- and
closed-loop proportional or servovalve circuits. The difference between the air valves in
these circuits is how fast they respond. Most proportional valves have a sealed spool that
controls direction and flow so the valves tend to hang up and jump. Pneumatic
servovalves often have spools with metal-to-metal fits that float on bypass air.

The valve shown in Figure 12-8 is sold as a proportional or servo directional control
valve for hydraulic or air circuits. Controlling the amount of air and which direction it
goes is not a problem but the compressibility of air creates some giant hurdles to
overcome.

Proportional valves usually control only acceleration, deceleration and/or speed because
these circuits do not include feedback transducers. It is very easy to get smooth
acceleration and deceleration with high speed in between without other controls or
shock absorbers to stop the load mechanically.

Adding feedback transducers to a proportional air circuit can provide servo-like control
for light loads -- such as those found in pick-and-place applications. However, a
proportional valve is usually not responsive enough for exacting part placement or
speed control. For very accurate control, a servovalve with feedback transducers can
give close-tolerance positioning (with light loads), repeatable velocity control, and very
accurate holding force.

Pneumatic proportional and servovalves are not a replacement for electromechanical or


servo hydraulics, but they have price advantages over both systems. When the loads are
light and cost is a factor, they are worth a look.

General information for hydraulic infinitely variable valves

 The symbol for proportional and servovalves shows a 4-way, 3-position function and the valve
can move to each of the positions. However, the parallel lines along the sides of the symbol
indicate the valve does not have to shift all the way all at once. These valves can shift into
straight or crossed arrows in any proportion from 0 to 100%. They are infinitely variable and
can pass any flow desired.
 servovalves are always 3-position, all-ports-blocked center condition, as shown by the symbols
in Figures 12-8 through 12-11.
 always size proportional or servovalves for high pressure drop. Proportional valves should have
200- to 500-psi pressure drop at full flow. Most servovalve manufacturers rate their valves at
1000-psi pressure drop at full flow. This means the valve may look physically small for a given
flow in relation to conventional valves. It also means most servo and proportional valve circuits
require a heat exchanger to deal with excess wasted energy.
 always mount the valve as close as possible to the actuator ports. Any piping between the valve
and the actuator holds extra fluid that can make the system softer and less responsive. This is
especially important on air-powered circuits.
 never use hose between the valve and the actuator. If isolation is necessary, mount the valve on
the moving part and use flexible lines for supply and return.
 use in-line pressure filters at the supply to each servovalve or bank of valves to protect the valve
from contamination in the pump and piping.

Specific Information for pneumatic infinitely variable valves

 use air at the highest pressure possible that does not exceed component or plumbing limits.
This is usually as high as 250 psi.
 size the valve to flow just enough to produce the maximum desired actuator speed. Oversize
valves produce erratic control because a small spool movement gives more flow than required.
 use an actuator with the largest area practicable so the load moves with a low pressure
difference. Note: the larger the cylinder, the more air it consumes so operating cost escalates.
 pneumatic servo circuits do not work well when outside forces push against the actuator. The
actuator tries to resist, but force buildup is slow in comparison to electromechanical or
electrohydraulic systems.

Typical servo circuits

Figure 12-14 shows schematic drawings of three typical servo circuits. In the figure,
each type circuit controls a different actuator, but any actuator could have more than
one type of control.
Fig. 12-14. Servovalve
and closed-loop electronic circuit for accurate position, force, and speed control
The typical power unit for a servo system is a pressure-compensated pump with an
accumulator or accumulators. A servovalve must always have a ready supply of fluid
because no matter how fast it reacts, without an immediate supply of fluid the system
will be sluggish. The pressure-compensated pump may be at full pressure when no
actuators are moving, but its flow is zero. Adding accumulators assures that there is no
wait for the pump to come on stroke before the actuator receives flow. (A full discussion
of how accumulators work and how they are applied is in Chapter 16.)
Again, pressure filters in the lines to the servovalves make sure they receive clean oil.
One filter could be sufficient for multiple valves when the valves are close to each other.
The reason for pressure filters in the valve lines is the pump constantly produces
contamination particles that will shut down flapper-type servovalves.

Cylinder 1 is in a position loop. It can be placed at a precise location repeatedly within


±0.0005 in. A programmable logic controller (PLC) sends a signal to the summing
amplifier’s control card. The signal passes on to the valve-driver card and then to the
valve coils. This signal shifts the servovalve to start cylinder movement at a set rate. The
linear potentiometer sends position feedback to the summing amplifier and modifies
valve position to find and maintain a certain position. Often, position control is paired
with speed control to accelerate the actuator to a certain speed, then decelerate and stop
it at the desired position.

Cylinder 2 is in a force loop. A certain size cylinder operating at a given pressure


produces a given force. This force can be calculated by multiplying area times pressure,
but the result is not exact. Friction from seals and between external machine members
can reduce this force by a few pounds or more on an operating machine. When an exact
force calculation is required, a servovalve-controlled cylinder that has a load cell for
feedback can keep forces within 1/2% with ease. The summing amplifier sends a signal
from the PLC and feedback from the load cell modifies the valve position to exactly
match the input signal to generate the desired force.

The hydraulic motor is in a speed loop that maintains the motor’s rpm when the fluid
viscosity, pressure, or load changes. A rotary device called an encoder constantly sends
rpm information back to the summing amplifier to open or close the servovalve as
needed. Just as a cylinder does, a hydraulic motor will slow when the load increases,
when fluid gets thinner due to temperature increases, or when system pressure
fluctuates as other actuators move. If the encoder sends a reduced-rpm signal back, the
servovalve opens to let more fluid in. If the hydraulic motor tries to speed up, the
servovalve closes enough to maintain the set speed.

Other infinitely variable valve applications

Proportional and servovalves can be applied to variable-volume pump controls to


accurately set flow from the pump in relation to feedback from an actuator. Most
pressure-compensated pumps available today have this feature as an option.

An example might be a closed-loop hydrostatic pump and motor that must stay at a
constant speed. Feedback from an encoder signals the pump to increase or decrease flow
as the motor overspeeds or underspeeds.

Proportional coils make excellent infinitely variable flow controls that can be controlled
electrically. They can act as simple throttle valves or perform as full-blown pressure-
compensated types that adjust to pressure fluctuations.

Common circuits for proportional flow controls include feed-speed changes to deal with
load fluctuations; hydraulic motor rpm changes to control product backup; or
maintaining constant speed as loads vary. (Pressure-compensated proportional flow
control valves are covered in Chapter 13.)

Proportional coils also make excellent infinitely variable pressure-control valves.


Pressure settings from remote locations or a PLC at anytime from minimum to
maximum almost instantaneously. Most of these pressure controls are for relief and
reducing functions but could be used on any of the pressure controls. (Proportional
pressure-relief valves were covered in Chapter 9.)
RELATED HAPTER 20: Brain Teasers

Table of Contents

Bud Trinkel | Oct 23, 2007

BRAIN TEASERS

Problem #1
Click here for answer to problem #1

Problem #2

Click here for answer to problem #2


Problem #3

Click here for answer to problem #3

Problem #4
Click here for answer to problem #4

Problem #5
Click here for answer to problem #5
Problem #6

Click here for answer to problem #6


Problem #7
Click here for answer to problem #7

Problem #8
Click here for answer to problem #8

Problem #9

Click here for answer to problem #9


Problem #10

Click here for answer to problem #10

Problem #11
Click here for answer to problem #11

Problem #12
Click here for answer to problem #12

Problem #13
Click here for answer to problem #13

RELATED

You might also like