You are on page 1of 24

Top of Form

Bottom of Form

Top of Form

Bottom of Form

Top of Form

Bottom of Form

Contamination control involves preventing contaminants


from entering a hydraulic system and placing filters in
strategic locations throughout the system to trap any
contaminants that do find their way into the fluid. But for
critical equipment, a successful contamination control
program also must include regular assessment of the
hydraulic fluid's cleanliness. Often, this must be done
every two to six months or after every 500 or 1000 hours
of operation, depending on the equipment's duty cycle,
operating environment, and how critical it is to overall
operation. Fig. 1. Portable contamination detection instruments
make it quick and easy to assess the cleanliness of
Experts also recommend that fluid be tested immediately hydraulic fluid, whether the equipment is in an
after any maintenance event that exposes the hydraulic industrial plant or at a job site.
system to the external environment. This could be when
a hose or other component is replaced or fluid is added
to the reservoir. Fluid replenishment can be particularly
troublesome because new fluid is notorious for being
dirty - often from improper storage and handling
practices.

Test labs or do-it-yourself?


Before the advent of portable contamination detection instruments, Figure 1, fluid testing was conducted for only the
most critical equipment and sent to a laboratory for analysis. This is still the most practical route for companies that do
not require fluid testing often enough to justify purchasing their own diagnostic equipment and training their personnel.
And even if a company has its own equipment, test labs still prove valuable for running multiple tests, interpreting
results, troubleshooting, and recommending appropriate action.
Test labs should also be consulted for periodic chemical analysis of hydraulic fluid. Even if contamination has been
brought to within acceptable limits, certain contaminants can alter a fluid's chemical composition (primarily the additives)
and render it ineffective. For example, additive depletion, over time, can reduce a fluid's lubricity, oxidation resistance,
or anti-foaming characteristics.
Excessive water and overheating are two conditions that can upset a fluid's chemical balance relatively quickly.
Therefore, if excessive water is found in a fluid or a system overheats, experts recommend not only finding and
correcting the source of the problem, but conducting chemical analysis of the fluid as well. Even if these problems do
not occur, hydraulic fluid suppliers generally recommend having fluid analyzed chemically at regular intervals - such as
annually - to identify potential problems and prevent them from occurring.
Portable particle counters and other diagnostic equipment have made it easy and convenient to monitor fluid cleanliness
of even non-critical equipment. In fact, many companies that have invested in their own particle counters and other
instruments monitor cleanliness of more equipment more often.
The higher reliability that results from this more-intense preventive maintenance adds to the return on their investment.
Furthermore, advanced techniques are being developed to make fluid-monitoring instruments even more sophisticated.
Equipment currently is under development to continuously monitor the condition of hydraulic fluid while equipment is
running.

Assessment starts with a sample


Regardless of the details of any contamination-monitoring program, its usefulness will depend on fluid sampling. Fluid
samples must accurately represent the condition of the fluid within the hydraulic system. This means that the sampling
technique and devices, as well as the container, must not contaminate the fluid sample.
The point from which samples are extracted should be determined by the information desired from the sample. For
example, sampling fluid from a system's pump discharge line will likely produce different results from a sample taken
from a return line. Fluid from the pump discharge line is more likely to contain pump wear debris than fluid from a return
line because filters would have captured pump wear debris before it would reach a return line.
Experts advise, however, that samples taken from the reservoir usually are the most unreliable. First, because a
reservoir acts as a storage device, its contents have accumulated over a relatively long interval, whereas fluid from a
hydraulic line is more representative of conditions at the time the sample is taken. Second, most reservoirs are
designed to minimize turbulent flow so contaminants can settle to the bottom and air can rise to the top. This makes it
difficult to obtain a sample with a representative concentration of water and other contaminants.

Fluid analysis techniques


Once fluid samples have been obtained, any of several methods can be used to analyze the size, concentration, and
nature of the contaminants. The most common analysis techniques for hydraulic systems are:
• particle distribution
• gravimetric
• ferrographic wear debris
• proton induced X-ray, and
• water content.
Each of these tests produces different results according to the type of information desired. Therefore, they should not be
viewed as competing technologies. Rather, the more tests that are conducted on a sample, the more knowledge that
can be gained. But no matter which technique is employed, obtaining a pure and representative sample is essential to
achieving accurate results.
Particle distribution summarizes the number of contaminant particles classified by size for a sample. Automatic
particle counters have gained wide acceptance for this previously time-consuming, laborious task that produced
inconsistent results. The widespread use of particle counters is a testament to their ease of use and consistent
reliability. They are often used by technicians at manufacturing and maintenance facilities.
Gravimetric analysis summarizes the total mass of solid particles above a given size for a specific volume of fluid.
Results are reported as mass density, usually mg/l. Unlike particle counting, gravimetric analysis quantifies only solid
particles, not water.
However, gravimetric analysis gives no indication of size distribution. So a sample may contain 25 mg/l of solid particles
greater than, say, 5 µm. But this gives no indication as to what percentage of particles are greater than 10 µm and how
many are greater than 15 or even 25 µm. As with particle counters, gravimetric analysis instruments often are used by
technicians to monitor contaminants in hydraulic systems.
Ferrographic wear debris analysis quantifies wear debris (primarily metals) in a fluid sample. Because the most
highly stressed wearing parts of of machine components are made of steel, wear debris usually are influenced by
magnetic fields. Ferrographic analysis can be used to evaluate a system's wear mechanisms, assess the severity of
wear, and identify the predominant materials being worn away.
Proton-induced X-ray emission (PIXE) summarizes the elemental composition of solid contaminants and wear debris
in a fluid. The procedure involves exposing the fluid sample to a proton beam. A computer then interprets results of the
test by producing data on the entire spectrum of elements in the target, not just a single element.
Neither particle counting nor gravimetric techniques can differentiate between foreign contaminants and wear debris,
which makes PIXE useful for gaining insight into the nature of particles found in a fluid.
Water-content analysis determines how much water is present in a base fluid. Next to particulate matter, water, by far,
is the most damaging contaminant in a hydraulic system - or any oil-lubricated system for that matter. Higher
concentrations of water in hydraulic oil accelerate wear, fluid degradation, corrosion, and reduction in service life.
Therefore, once the amount of water present in a hydraulic fluid has been found, the challenge becomes determining
how much can be tolerated.
The test itself uses a solution that conducts electrical current based, in part, on the amount of water contained in a
sample. Measuring the current and its duration provides an indication of water content in the base fluid. The test can be
accurate to within 10 ppm, but additive packages common to hydraulic fluid tend to produce less detailed results.

New technology enhances assessment


Whether your car has an "idiot light" or a real temperature gauge, its designers deemed the engine's water temperature
important enough to monitor it continuously. After all, it wouldn't do much good to have your engine's water temperature
checked only when you stop for gas. It's much more likely that a problem would occur while you were on the road rather
than when you were at the gas station.
Monitoring contamination of hydraulic fluid would at first seem to be much less critical than engine water temperature.
After all, fluid usually becomes contaminated gradually, so monitoring its condition frequently enough can identify
problems before they cause any real harm. Engine temperature, however, can increase quickly once a problem occurs.
If a hose ruptures, the water pump gives out, or the radiator leaks, the engine can quickly overheat.
Contamination, under certain conditions, can also act quickly to cause catastrophic failure in a hydraulic system. For
example, if a pump ingests enough air to cause serious cavitation, it can become inoperative within days. Or if a large
quantity of water flows through a system, hydraulic fluid can lose its lubricity, which will result in rapid wear of
components. If either of these events occurred a few weeks before a scheduled fluid analysis, the machine could
undergo costly downtime.
Granted, these types of problems happen rarely. But if the equipment costs millions of dollars or works in an operation
where downtime is measured in thousands of dollars per hour, it becomes practical to continuously monitor fluid
cleanliness. It is for cases like these that companies are developing systems to monitor the cleanliness of hydraulic fluid
continuously while a system is running.
One such prototype system routes pressurized fluid from the pump into a tube through which light is transmitted. When
the fluid is clean and relatively free of air and water, a receptor detects the amount and pattern of light transmitted
through the fluid.
As the fluid becomes more contaminated, the amount and diffraction of light transmitted through the tube changes. If
undissolved water or air is present, the transmitted light becomes more scattered. Calibrating the receptor to these
different conditions provides an instantaneous indication of the fluids condition. Therefore, a potentially catastrophic
failure can be averted by taking appropriate action immediately.
Another emerging technology is a system that enables users to go beyond particle counting and actually analyze wear
debris. The system consists of hardware to generate digital photomicrographs and software to aid in analyzing the
digital images. The hardware includes a digital camera, an adapter to mount the camera to a microscope, and a PC-
compatible image-capture card.
Once imported to a PC, images can be compared to those in an atlas of known wear debris using wear debris analysis
software. The software also aids in characterizing descriptions, managing data, and generating reports. The analysis
can also be incorporated into maintenance and SPC software used for plant operation and quality assessment.

Special fittings help keep samples clean

Tapping into a hydraulic system to sample fluid creates the potential to


contaminate not only the fluid in the hydraulic system, but the fluid sample
as well. To help prevent either from occurring, test ports designed for fluid
sampling should be mounted permanently to the equipment and have
protective caps to keep dirt away from the sampling port. The cap is
removed only when taking a fluid sample and replaced immediately
afterward.
Shown at right is a low-pressure test fitting for sampling fluid from
hydraulic return lines without having to shut down equipment. Pressing on
the pushbutton opens a check valve that routes fluid from the hydraulic
system out through the sampling port. Models for taking samples from
high-pressure lines work in a similar fashion, but use a threaded Special fittings are available to
connection instead of a pushbutton to open the check valve. minimize the potential for
contamination when sampling fluid
The test port avoids introducing external contaminants into the fluid from a hydraulic system. Shown here is
extracted from the hydraulic system. However, to ensure accurate a model for sampling fluid from a
samples, tubing leading to the sampling vessel and the sampling vessel hydraulic system's return line at
itself must be absolutely clean. The tube should be discarded after a pressures to 600 psi. It comes in a
sample is drawn and replaced with a new one before each subsequent variety of port sizes and thread
sampling. configurations.
Don't ignore compressed-air cleanliness
Manufacturers of compressed air components often cite wet or dirty air as the cause of problems related to a machine's
operation or performance. After all, air is a fluid, and accumulated contaminants tend to restrict or even block flow. And
with many of today's pneumatic components requiring increasingly cleaner air, the potential consequences of
contamination are becoming more widespread and pronounced.
Therefore, it is more important than ever that pneumatic systems contain air drying equipment and properly rated filters.
Furthermore, as with hydraulic fluid, compressed air cleanliness should be assessed periodically to ensure maximum
performance and reliability. This task has become easier thanks to portable instruments that can monitor compressed air
contamination concentration, pressure dewpoint, air temperature, and pressure.

Tolerances that exist in today's highpressure hydraulic systems demand tight control of system contamination.
Contamination that is built into systems during manufacture and assembly must be removed before start-up to insure
proper and predictable system performance throughout its service life.
A new or rebuilt hydraulic system should be flushed before it becomes operational. The concept of flushing is to loosen
and remove contamination particles inside the system by forcing flushing fluid through it at high velocity. In theory, this
leaves the inside walls of the fluid conductors at the same cleanliness level as the new fluid that will be installed. Then,
during normal operation, the system will experience only externally and internally generated contamination that can be
controlled with filtration.
Instructions for flushing usually specify a level of system cleanliness that must be achieved, and sometimes a fluid
velocity that must be maintained during the flushing procedure. Typical instructions state that flushing must be
accomplished at normal system fluid velocities for a certain period of time with a certain level of filtration. More stringent
specifications may call for a particular fluid contamination level and require documentation by fluid-contamination
analysis.
One shortcoming of all these flushing methods is that they are based on procedures to clean the fluid, but ignore the
interior cleanliness of the system. Even if the tubing and conductors have been installed with the greatest of visual care,
the human eye can only see particles that are larger than 40 µm - well below the needs of even the crudest and most
elementary hydraulic system.

How high a velocity?


The critical variable in flushing to achieve acceptable fluid and conductor cleanliness is fluid velocity. Traditional flushing
methods usually establish this velocity in one of two ways:
• the velocity must be high enough to achieve a Reynolds Number (NR) of 3,000 or more, or
• the velocity must meet or exceed the system fluid's normal operating velocity as designed.
Experience has shown that neither of these flushing velocities is sufficient to assure the cleanliness of the ID of the
system's conductors. A short review of basic fluid dynamics explains why.
Dimensionless NRs are used (along with other factors) to classify fluid flow as either laminar, turbulent, or somewhere in
between, Figure 1. NR depends on the fluid's viscosity and velocity and the ID of the pipe. The flow condition that exists
when NR is less than 2,000 is termed laminar, signifying orderly flow with parallel streamlines.1 When NR is greater than
3,000, the flow becomes turbulent, defined as the condition when fluid streamlines are no longer orderly. Between
Reynolds Numbers of 2,000 and 3,000, flow exists in transition, sometimes called the critical zone.
The hydraulic fluid velocity required to achieve turbulent flow is well within the recommended fluid velocity guidelines for
hydraulic fluid conductors.2This equation reinforces that statement:
NR = VD/v
where:
V is the fluid velocity in ft/sec,
D is the ID of the fluid conductor in ft, and
v is the fluid kinematic viscosity in ft2/sec.

First example
Suppose that the NR is 3,000, the conductor is a 1-in. tube with a wall thickness of 0.049 in., and v is 1.288X10-4 ft2/sec.
Calculated fluid velocity then is 5.14 ft/sec, which corresponds to a flow rate of 10.24 gpm.
Note that the viscosity (and therefore the Reynolds Number) of a typical hydraulic fluid is influenced by temperature and
pressure. That is, the hotter the oil, the higher the NR for the same fluid velocity and pressure. The higher the pressure,
the lower the NR for the same fluid velocity and temperature. Thus, specifying that NR should be 3,000 is not a stringent
requirement, but is well within the normal operating fluid velocities of a system. By definition, turbulent flow has been
created because the fluid streamlines are no longer parallel, but sufficient fluid motion to clean the inside walls of the
conductors has not been generated.
Even at the recommended maximum fluid velocities and NRs for hydraulic-system working conductors, fluid flow still is
not turbulent enough to greatly affect contamination on conductor walls. Boundary- layer fluid at the interior surfaces of
the fluid conductor remains undisturbed.
Second example
The NR for flow at normal system velocities next can be calculated using the same conductor size and kinematic
viscosity as in the first example, but with the velocity increased to 20 ft/sec. This higher velocity results in a Reynolds
Number of 11,671, which corresponds to a flow rate of 39.8 gpm.
As NR increases, flow conditions go from laminar, through the critical zone, to turbulent. It has been proven empirically
that once NR exceeds 3,000, resistance to fluid flow is a combination of the effects of turbulence and of viscous drag at
the conductor wall. (This region of viscous drag at the conductor wall is known as the viscous sub-layer.) There is a
transition zone within the turbulent flow range where flow resistance goes from being governed by turbulence effects to
being governed by the roughness of the inside wall of the conductor.
This is shown clearly when inspecting the Moody diagram,3 Figure 2, which graphically demonstrates the relationship
between Reynolds Number NR, friction factor f, and e, the roughness of the conductor's inside surface. Resistance to
flow through a fluid conductor, represented by friction factor f, is only affected by the surface roughness of the fluid
conductor4 when NR is above 4,000. Thus, the majority of the resistance to flow is created by turbulence effects. Only
when NR is high enough so that surface projections of the conductor walls extend beyond the viscous sub-layer does
the surface come in contact with the turbulent flow and affect the pressure drop in the conductor.

Surface roughness

Fig. 2. Modified Moody diagram relates friction factor f, Reynolds Number NR, and conductor surface roughness e.

For drawn tubing, average surface roughness e is 0.000005 ft. If the conductor is the same 1-in. tubing with 0.049-in.
wall thickness, ratio e/D will be 0.000067. The Moody diagram indicates that for this conductor, NR must be at least
25,000 before the inside surface exposes its resistance to fluid flow. To ensure the inside wall of the conductor will be
cleaned,NR must be greater than 25,000. For flow to be fully in the rough zone of turbulent flow, NR must be greater than
3.25X107. Using 1.288X10-4 ft2/sec, the same fluid kinematic viscosity as in the first example, a NR of 25,000
corresponds to a fluid velocity of 42.8 ft/sec, or a flow rate of 85 gpm - still easily attainable with conventional hydraulic
pumps.

Real-world systems
It can be argued that if the walls of a conductor are not greatly affected by normal system fluid velocities, contaminants
lodged there will have little chance of entering the fluid stream. This may be partially true but the argument applies only
to smooth, straight conductors at steady flows and pressures. It is not representative of normal installations that
combine straight runs, bends, and numerous fittings where flow patterns are only predictable empirically, and where
pressure fluctuations and spikes are commonplace.
Depending on the severity of service that the system will experience, pressure spikes will dislodge contaminants held in
the walls of the conductors and between fitting interfaces. In critical systems, 3- to 25-m particles can impact system
performance significantly. The only way to guarantee that conductor contamination (that can be released at any time
during operation) does not affect system performance is to protect each component with a filter, an option so costly that
it would not be used in most systems. Although flushing hydraulic system conductors at the normal system
operatingfluid velocities can provide fluid velocities higher than flushing at a NR of 3,000, the inside wall of the
conductors still will not be cleaned.
High-velocity/high-pressure flushing
Flows that produce NRs .25,000 are needed to ensure that conductor walls are exposed to turbulent flow. Because
system conductors may consist of pipe, tube, and/or hose and associated fittings, the specification of a contractual
number for NR is difficult and still does not guarantee that conductors will be cleaned. The best one can do is establish
conditions that will maximize NR. These conditions are: the highest possible velocity at the lowest possible fluid
viscosity. Limiting factors are the conductor's pressure rating and the fluid's maximum operating temperature.
When flushing a system, the valving and actuators must be "jumpered" for safety reasons so the only resistance to fluid
flow is the pressure drop in the conductors and fittings. When flow becomes turbulent, the pressure drop is proportional
to the square of the velocity. Extrapolating this relationship to its maximum, the highest possible velocity occurs when
the pressure drop in the conductor generated by fluid flow is equal to the maximum test pressure of the conductor.
Flushing a system at these high flows and pressures has the added advantage of expanding and contracting the
conductors and fittings as the pressure fluctuates while inducing highly turbulent flow. This optimizes the flushing action.
By equating the pressure drop in a conductor to the maximum pressure rating of that conductor, the maximum fluid
velocity possible, along with the corresponding NR, can be calculated. The temperature of the fluid directly affects its
viscosity and is the other variable that can control NR. Flushing pressure also affects viscosity, but this is hard to
quantify because pressure in the pipe being flushed will vary from maximum at the pumping source to atmospheric at
the conductor outlet.
The equation used to calculate head loss in the turbulent zone is:
hl = fLV2/2D ,
where:
hl stands for head loss,
f is the friction factor found in the Moody diagram,
L is the conductor length in ft,
V is the fluid velocity, and
D is the conductor's ID in in.
This equation will calculate the maximum velocities and NRs that can be achieved for any given maximum flushing
pressure.
Determining f for pipe flow requires iterative calculations using the Moody diagram. Given the pressure rating, ID,
length, and relative roughness of the conductor, assume an f and then calculate the fluid velocity. next calculate NR and
determine a new f from the Moody diagram. Repeat the calculation until f converges.

The table above contains velocities and NRs that have been calculated for 200 feet of Schedule-80 pipe using the
maximum test pressure for the pipe and a surface roughness of 0.00015 ft for wrought iron pipe. These calculations did
not take into account the pressure drop produced by the various fittings normally used, so the values for the attainable
fluid velocities and NRs are optimistically high. Also, special fluids with lower viscosities or flushing at higher
temperatures to reduce the fluid viscosity can increase NR.
The values determined for maximum flushing velocity and flow rate indicate that some of these conditions - mainly for
lines with inside diameters smaller than 3/4 in. - can be satisfied using conventional high-pressure pumps of appropriate
flow capacity, although it may be difficult to induce the pressure fluctuations needed to dislodge contaminants. For
systems with larger conductors, special methods must be used to achieve the necessary pressures, fluid velocities, and
NRs to properly flush the lines.

Clean hose prevents contamination troubles


Research has proven that clean oil and a clean
hydraulic system are the keys to longer system life
and less downtime. However, clean oil can be
contaminated by particles left behind from the cutting,
crimping, bending, and flaring process, creating a real
problem. Attempts have been made to solve this
problem by using chemicals, high pressure flushing,
and pull through techniques with marginal results.
A relatively new cleaning technique offers an effective
and practical alternative to these procedures by using
sponge-like projectiles that are shot through hose and
tubing assemblies by a blast of compressed air. Equipment for this procedure is much more affordable than that for
high-velocity flushing. Perhaps more importantly, though, it cleans assemblies much more effectively than compressed
air alone and in a fraction of the time it takes for high-velocity flushing.
Launchers fire the projectile (which is sized 20% to 30% larger than the I.D. of the hose, tube or pipe), stripping out the
internal contamination as it travels through couplings and around bends - forcing the contamination out in front of it.
Once the hose assembly has been cleaned, be sure to install protective caps or plugs at both ends to prevent
contamination from entering the assembly. These should not be removed until the hose assembly is being installed on
the equipment.

Concern for the environment has become a fact of life for most companies. For the companies that use hydraulic
equipment, that concern is reinforced by stringent federal and local regulations requiring pollution prevention and
governing disposal of used industrial fluids. Higher costs (and greater potential liabilities) for fluid disposal are a direct
result of these regulations. This adds another economic item to the list of benefits derived from controlling contamination
to extend the service life of hydraulic fluids. The list now includes:
• lower cost for fluid needed for replacement and replenishment
• more consistent hydraulic system performance
• less component wear, and
• reduced fluid disposal costs.

Contamination's detrimental effects


It is an established fact that particulate contamination and water in hydraulic fluids can have serious adverse effects on
the fluids' physical and chemical properties. The loss of crucial fluid properties, which are central to useful service life,
can result in inefficient system performance and accelerated mechanical and chemical wear processes.
Hydraulic fluids are carefully formulated for specific areas of application. They usually are comprised of a base stock
and an additive package. The additive package consists of chemical compounds designed to protect the base stock - as
well as the components in the hydraulic system - and to ensure proper performance of the system. Typical additives
include dispersants and detergents; anti-oxidants; anti-corrosion, anti-wear, anti-foaming and extreme pressure (EP)
agents; and viscosity-index improvers. Particulate contamination and water adversely affect both the base stock and the
additives.
Water is a poor lubricant, and significant concentrations of water in hydraulic fluids can decrease their viscosity and
load-carrying ability, as well as hydrodynamic-film thickness. This can lead to greater surface-to-surface contact at
sliding and rolling dynamic clearances, and hence, increased component wear. The presence of free water in systems
that could be exposed to temperatures below the freezing point of water can lead to icing, which will degrade system
performance and can produce malfunctions.
The presence of both water and particulate contamination can lead to the formation of insoluble precipitates, and
viscous sludges and gels. These materials induce excessive stress on system components - especially pumps - and
can clog orifices, nozzles, and jets.

Degradation of fluid base stock


Oxidation of the fluid base stock is a primary chemical-degradation process in many hydraulic fluids. The oxidation
process proceeds through a series of chemical chain reactions and is self-propagating - with the intermediate, reactive
chemical species regenerating themselves during the process. The result is the formation of oxygenated compounds
(notably acidic compounds in the case of hydrocarbon, polyol-ester, or phosphate-ester base stocks) and, eventually,
high-molecular-weight polymeric compounds. All of these compounds often are insoluble; they settle out of the fluid as
gums, resins, or sludges.
Oxidation is significantly accelerated in the presence of metals and water. Metals act as catalysts, and fine metallic wear
debris, commonly found in hydraulic systems, are especially active, due to their large effective surface area.
Table 1 (below) summarizes data from tests that were carried out to quantify the effect of metal catalysts and water on
oil oxidation. The tests were conducted on turbine-grade oil in a pure oxygen atmosphere, according to the ASTM/D-
943 oxidation test procedure. The neutralization number, tabulated in the last column, is a measure of the extent of
oxidation. The results show that the extent of oxidation is greatly increased: roughly 48-fold for iron/water and 65-fold for
copper/water within 400 and 100 hours, respectively - compared to the baseline test with no water and metal catalysts.
Even with only a single contaminant present (either water or a metallic catalyst), the neutralization numbers increase.
Hydrolysis - the alteration or decomposition of a chemical substance due to the presence of water - is another potential
problem in hydraulic fluids. Fluid base stocks that are comprised of ester compounds, such as polyol esters and
phosphate esters, can undergo hydrolysis under typical hydraulic-system operating conditions. The acidic compounds
that may form during hydrolysis can react with materials of the hydraulic-system components, leading to corrosion and
insoluble corrosion products.

Additive depletion
Depletion of additives can occur either by their physical removal from the fluid or by chemical reactions which convert
them to non-functional products. The solubility of many additives is critically dependent on fluid composition. The
presence of water can lead to the precipitation of these additives from the fluid. In addition to being rendered non-
functional, the precipitated additives contribute to the particulate contamination level in the fluid.
Additives that protect the base stock can be depleted rapidly due to the enhanced degradation of the base stock in the
presence of both particulate contamination and water. A notable example is the depletion of antioxidants. A summary of
the detrimental effects of particulate contamination and water is presented in Table 2. (below)

Monitoring contaminants
How can you check on the status of the hydraulic fluid in an active system? Not too long ago, you had to take samples
of fluid from the system, send them to a laboratory, and wait for a report. Today, a variety of on-line monitors is available
for in-house measurement of particulate levels and water concentration. Some monitors can be interfaced with fluid-
purification devices for automated operation. When the monitor detects that preset threshold concentration limits for
particulates and/or water have been reached, the contaminated fluid is directed into the purifier - with no operator
intervention required.
On-line particle counters or contamination monitors are used most commonly to measure particulate levels in fluids in
installed systems. Particle counters provide an estimate of the particle size distribution, i.e., particle size vs number of
particles, for several preset size ranges (usually six to seven). They operate on the principle of light obscuration by
particles in the fluid-sample stream. Portable models, Figure 1, can be moved to multiple locations or installed at a
single location for continuous on-line sampling. The contamination level typically is quantified in terms of the fluid
cleanliness code (ISO 4406) or fluid cleanliness classes (NAS 1638 or ISO 11218).
Contamination monitors operate on the principle of mesh blockage by particles in the fluid-sample stream. They provide
an estimate of the fluid cleanliness level in terms of cleanliness codes or cleanliness classes. Contamination monitors
are the preferred choice for systems where the fluid properties prevent the use of light-obscuration particle counters:
dark fluids that transmit light poorly or fluids containing air bubbles or emulsions.
Typical on-line water monitors, Figure 2, are adaptations of devices that measure relative humidity. They indicate the
percent saturation of water in the fluid - free water is 100% saturation - and the temperature. Note that the saturation
level for water in hydraulic fluids depends on the specific fluid composition (both base stock and additive package), the
actual condition of the fluid, and its temperature. The same type fluid, formulated by different manufacturers, could differ
in water saturation levels. Likewise, the saturation level of the same fluid could differ over a period of time. Thus,
correlation of the absolute water concentration, i.e., ppm concentration, with percent saturation, requires determination
of the absolute water concentration in the specific fluid in question. In view of the above, it is most convenient to specify
water concentration limits in terms of percent saturation.

Removal of water and particulate


Several methods also are available to remove particulate contamination and water from hydraulic fluids. The choice of
method depends both on the contamination level of the fluid and its specific area of application. Heavily contaminated
fluids are best cleaned by removing them from the operating system and purifying them externally prior to re-use.
Subsequently, in-line particulate filters and water-absorbing filters can provide contamination control.
Although a variety of equipment is available to remove free water (e.g. centrifuges, coalescers, and water-removal
cartridges), only fluid purifiers offer the ability to remove free, emulsified, and dissolved water. In addition to removing
water and particulate contamination, fluid purifiers also take out volatile solvents and dissolved gases.
Two common types of purification devices are flash-distillation and vacuum-dehydration systems. In flash-distillation
systems, fluid is heated and then introduced into a vacuum chamber so that free and dissolved water, gases, and
solvents are distilled off, thus dehydrating the fluid. In vacuum-dehydration systems, the fluid is exposed to a low-
humidity atmosphere in a partial vacuum chamber, resulting in the transfer of free and dissolved water, solvents, and
gases from the fluid to the atmosphere in the vacuum chamber. To facilitate the transfer, the surface area of the fluid
should be maximized.
In one purifier design, Figure 3, the contaminated fluid is introduced into the vacuum chamber through fine spray
nozzles to form a conical, thin film through which a flow of low-humidity air is directed. This arrangement results in a
large fluid-surface area that allows for more efficient transfer of water from the fluid film to the air stream. The air then is
exhausted through a de-mister filter to remove any residual fluid it may be carrying.
In the final stage of the purifier, de-aerated and dehydrated fluid exits the vacuum chamber through a particulate
contamination-control filter. These fine filters have high efficiency particle-removal characteristics, especially in the
smaller size ranges. For example, their particle-removal efficiency is 99.5% or higher for particles greater than 3 µm in
size, i.e., β3>200. These filters also exhibit a high dirt-retention capacity.

At work in the real world


One U.S. airline studied the impact of contamination on performance in the hydraulic systems of its aircraft ground-
support equipment, such as mobile cargo loaders, container rotators, aircraft bridges, and nose docks. This equipment
operates outdoors and is exposed to dirt and weather extremes. Initial investigations revealed high particulate levels -
20/16 on the ISO 4406 scale - and water contamination in excess of 1,000 ppm. These conditions required fluid changes
every two to three months. In spite of these relatively frequent changes, equipment failure was common.
The airline initiated a comprehensive program of contamination control. It included installing high-efficiency fine filtration
and the use of portable fluid purifiers. The result: hydraulic-fluid service life was extended to more than eleven months.
The water concentration in the fluid was held consistently below the manufacturer's recommended 200- to 400-ppm
level. Particulate levels were reduced to 13/11 on the ISO 4406 scale. The associated benefits were improved
performance and less downtime.
Dealing with acidic components
Phosphate-ester fluids are particularly susceptible to hydrolysis during service, resulting in an accumulation of acidic
hydrolysis products. A new trend in fluid purifiers is the incorporation of ion-exchange resin cartridges to remove acidic
components from phosphate-ester fluids. If acidic components are a problem, these purifiers provide a double benefit:
they remove water from the fluid to reduce the possibility of hydrolysis, and they adsorb any acidic hydrolysis products
that already exist in the fluid to minimize acid buildup.

Table 1. Effect of metal catalysts and water on oil oxidation


Final
Catalyst Water Hours neutralization
number
None No 3,500+ 0.17
None Yes 3,500+ 0.90
Iron No 3,500+ 0.65
Iron Yes 400 8.10
Copper No 3,000 0.89
Copper Yes 100 11.20

Table 2: Effect of particulate contamination and water on hydraulic fluids


Fluid breakdown Cause Effect on system
a. Agglomeration and precipitation of particulate  component wear
contamination  clogging of jets, nozzles, and orifices;
Physical
b. Oxidation/hydrolysis products - gums and sludges valve jamming
properties
c. Reactions involving additives - sludges and solids  system malfunction due to icing of free
d. Free water water
Base-stock a. Oxidation  corrosion and surface degradation of
degradation b. Hydrolysis components
a. Precipitation of additives
b. Adsorption by particulates  loss of component protection
Additive depletion
c. Reactions involving additives  increased component wear and corrosion
d. Abnormal degradation of base-stock
For several decades, a product called Air Cleaner Fine Test Dust (ACFTD) served as the standard solid-particle
contaminant for a number of purposes in the area of hydraulic contamination measurement and testing. The irregularly
shaped ACFTD particles - ranging in size from roughly 0 to100 mm - were very similar to the contaminants found in
typical hydraulic systems. In the particle-size distribution defined by ISO Standard 4402, ACFTD was used to set the
electronic threshold levels that establish the particle sizes measured in automatic particle counters (APCs). The dust
also was added to fluids in filter-performance testing to measure both the efficiency and dirt-holding capacity of filter
media. In addition, ACFTD was used to test the contaminant sensitivity of hydraulic components.
The AC Spark Plug Div. (later the AC Rochester Div.) of General Motors Corp. manufactured ACFTD by collecting dust
- primarily silica - from a certain area in Arizona, then ball milling and classifying it into a consistent particle-size
distribution. But in 1992, GM announced that it would discontinue production of ACFTD.
As a result, ISO Technical Committee TC 22 and the Society of Automotive Engineers (SAE) went to work to find
suitable test dusts to replace the old standard. Their efforts produced a new standard, ISO 12103-1, 1997, which
defines and designates four new test dusts. Powder Technology, Inc. (PTI), Burnsville, Minn., manufactures these dusts
from the same silica-based material used by AC Rochester so that their chemical characteristics are similar to the AC
Test Dusts. In a slightly different production method, PTI processes the Arizona dust with a jet mill and then classifies it.
Of the four new dusts, ISO Medium Test Dust (ISO MTD) has a particle-size distribution closest to ACFTD and,
therefore, has been selected as the replacement dust for APC calibration and filter-testing purposes.
While it is very similar to ACFTD, ISO MTD produces test results that are somewhat different. Therefore, results of both
automatic particle counting and laboratory filter performance testing (including filter efficiency and dirt-holding capacity)
can be significantly affected. Note that this is an artifact of the testing only; filter performance and actual contamination
levels in the field will remain the same as before.

NIST certification and new calibration


The US National Institute of Standards and Technology (NIST) undertook a project to certify the particle-size distribution
of ISO MTD suspensions in oil. From this study, the Institute determined that for particle sizes below 10 µm, the actual
particle size is greater than previously measured using an automatic particle counter that was calibrated with ACFTD.
Above 10 mm, the particles were smaller than under the old calibration system. To make a distinction, particle sizes
based on the new NIST determination will be represented as X µm(c), with the (c) designation referring to certified
calibration that has sizes traceable to NIST. Thus, particle size has a new definition, Table 1.
ISO Technical Committee TC 131 replaced the old APC-calibration procedure, ISO 4402, with a new procedure, ISO
11171, which incorporates ISO MTD test dust with the NIST particle-size and particle-count determinations. It also has
included a number of other enhancements to ensure better accuracy, reproducibility, and repeatability. In addition, ISO
has developed another procedure, ISO 11943, for calibration and verification of on-line automatic particle counters.
Because APCs are used for multi-pass filter performance testing and fluid contamination measurement, these changes
will affect reported results.
ISO has adopted a revised procedure for reporting fluid-cleanliness measurements from APCs that have been
calibrated with the new NIST-traceable method. This procedure, ISO 4406: 1999, uses three code numbers that
correspond to concentrations of particles larger than 4, 6, and 14 µm(c). The new 6- and 14 µm(c) sizes correspond
closely to the older 5- and 15-µm sizes reported by the old ISO 4406 coding system measured with an APC calibrated
with ACFTD. The new 4-µm(c) size, however, corresponds to about the 1-µm size if the ACFTD calibration procedure
had been used. This difference results in somewhat higher values of the first digit when compared to current 3-digit
codes that reference particles larger than 2 µm using the old ACFTD-calibration method.

Filter-performance testing
A number of substantial changes were made to ISO 4572, the multi-pass filter test procedure. These changes again
were intended to produce more repeatable and reproducible test results. The new method, ISO 16889 replaces ISO
4572, and incorporates ISO MTD and the NIST-traceable APC-calibration procedure. Beta ratios derived from tests that
use this new ISO procedure also are designated with the symbol (c) to signify they were measured in accordance with
the ISO 16889 procedure using NIST-traceable calibration. As an example, a beta ratio of 200 at 5 µm(c) would be
designated as β5(c) = 200.
The revisions to the multi-pass test method and the inclusion of both ISO MTD and the new APC-calibration procedure
will dramatically affect reported beta ratios for filter elements. The effect will vary for different filters, depending on the
influence of the test dust and the degree of change in the particle size at the filter's rating. In general, fine filters will
appear coarser or less efficient, and coarse filters will appear finer or more efficient.
We performed tests on Pall Industrial Hydraulics standard filter media using the new ISO 16889 test method.
Comparisons to results obtained from the previous ISO 4572 method are shown in Table 2. Note that filter performance
in the field does not change at all. The media are no more or less efficient at removing harmful particles. Only the
reported laboratory results have changed slightly because of the new procedures and methods.

Filter dirt-holding capacity


The replacement of ACFTD with ISO MTD in the multi-pass test also affects retained-dirt-capacity values for filter
elements. Capacity may be somewhat higher or lower with the new dust, again depending on the specific filter being
tested. However, most filters we evaluated exhibited an increase of about 10% to 40% in dirt-holding capacity when
using ISO MTD. Because each type of filter performs differently with the new dust, there is no single conversion factor
to change ACFTD capacities into ISO MTD capacities.
Again note that an increase or decrease in dirt capacity when tested with ISO MTD does not imply that the filter's actual
service life will be longer or shorter. In fact, there will be no change in field-service life. As a rule, dirt-holding capacity
should not be used as an indicator of field-service life.

Reporting fluid cleanliness


The changes discussed in the previous sections will not affect the actual cleanliness of fluids in the field. However, for
those technicians reporting data using particle counts, the changes in APC calibration will affect results obtained from
both laboratory and portable equipment that is calibrated to the new standards, as in Table 3.
Technicians reporting data using the ISO cleanliness codes will notice less effect. The adoption of a cleanliness code
with 3 digits should not have any appreciable impact because 3-digit codes have been used by industry for a number of
years. In addition, the change from 5- and 15-µm sizes to 6- and 14-µm(c) sizes will not show a significant change in
fluid-cleanliness code. The primary change will come from adoption of the 4-µm(c) size. This is a new addition to the
ISO 4406 standard (although a third digit at 2 µm using ACFTD had been used). Because the new 4-µm(c) size equates
to about 1 µm using ACFTD calibration, particle counts on fluid samples will typically show an increase of about one
level for this first digit of the code, Table 4. The reason: there usually are more smaller particles in any fluid sample.

Conclusion
ISO MTD was chosen as a replacement for Air Cleaner Fine Test Dust because ACFTD is no longer being
manufactured. To gain better resolution, accuracy, repeatability, and reproducibility in tests using the replacement dust,
four new or revised ISO standards have been adopted to accommodate the new test dust. These changes will have an
impact on automatic particle-counter calibration, particle-size definition, and laboratory reporting of filter performance -
both in particle-removal efficiency (fine filters will appear coarser and coarse filters finer) and dirt-holding capacity
(capacity will likely increase). The changes also will effect laboratory reporting on system fluid cleanliness (typically
showing a higher contamination level at smaller particle sizes).
This will undoubtedly cause some confusion, but remember that any impact is merely an artifact of the minor changes in
testing. Actual filter performance and field fluid contamination levels will remain the same.
Table 1. Measured particle sizes
Size in µm Size in µm(c)
with ACFTD NIST calibrated
(ISO 4402: 1991) (ISO 11171)

<1.0 4.0

1.0 4.2

2.0 4.6

2.7 5.0

3.0 5.1

4.3 6.0

5.0 6.4

7.0 7.7

10.0 9.8

12.0 11.3

15.0 13.6

15.5 14.0

20.0 17.5

25.0 21.2

30.0 24.9

40.0 31.7

50.0 38.2

100.0 70.0

Table 2: Comparison of laboratory filter ratings


Micron rating for Beta value
Pall media guide
Using ISO 4572 Using ISO 16889

βx=200 βx=1000 βx(c)=200 βx(c)=1000

KZ <1.0 1.0 2.0 2.5

KP 3.0 5.3 3.8 5.0

KN 6.0 8.3 5.7 7.0

KS 12.0 na 9.7 12.0


KT 25.0 na 18.2 22.0

Table 3: Typical effect of new calibration on particle counts


APC with old (ACFTD) calibration APC with new (NIST) calibration

Particle size in µm Particles/mL Particle size in µm(c) Particles/mL

2 4,170 2 24,900

5 1,870 5 3,400

15 179 15 105

25 40 25 14

Hydraulic fluids perform four basic Table 4: ISO code example


functions. Their primary function is to
create force and motion as flow is Old 2-digit ISO code
converted to pressure near the point (5 µm/15 µm)
of use. Second, by occupying the 14/12
space between metal surfaces, the
fluid forms a seal, which provides a
Old 3-digit ISO code
pressure barrier and helps exclude
contaminants. A third function - often (2 µm/5 µm/15 µm)
misunderstood - is lubrication of 16/14/12
metal surfaces. The fourth and final
function provided by hydraulic fluid is New 3-digit ISO code
cooling of system components. (4 µm(c)/6 µm(c)/14 µm(c))
If any one of these functions is 17/14/12
impaired, the hydraulic system will
not perform as designed. Worse yet,
sudden and catastrophic failure is possible. The resulting downtime can easily cost a large manufacturing plant
thousands of dollars an hour. Hydraulic-fluid maintenance helps prevent or reduce unplanned downtime. It is
accomplished through a continuous program to minimize and remove contaminants.
Aside from human interference, the most common source of system impairment is fluid contamination. Contamination
can exist as solid particles, water, air, or reactive chemicals. All impair fluid functions in one way or another.
Sources of contaminants

Contaminants enter a hydraulic system in a variety of ways. They may be:


 built-in during manufacturing and assembly processes

 internally generated during normal operation, and

 ingested from outside the system during normal operation.

If not properly flushed out, contaminants from manufacturing and assembly will be left in the system. These
contaminants include dust, welding slag, rubber particles from hoses and seals, sand from castings, and metal debris
from machined components. Also, when fluid is initially added to the system, a certain amount of contamination
probably comes with it. Typically, this contamination includes various kinds of dust particles and water.
During system operation, dust also enters through breather caps, imperfect seals, and any other openings. System
operation also generates internal contamination. This occurs as component wear debris and chemical byproducts from
fluid and additive breakdown due to heat or chemical reactions. Such materials then react with component surfaces to
create even more contaminants.
Contaminant interference
In broad terms, contaminant interference manifests itself as either mechanical or chemical interaction with components,
fluid, or fluid additives.
Mechanical interactions, Figure 1, include blockage of passages by hard or soft solid particles, and wear between
particles and component surfaces.
Chemical reactions include: formation of rust or other oxidation, conversion of the fluid into unwanted compounds,
depletion of additives - sometimes involving harmful byproducts - and production of biochemicals by microbes in the
fluid.
Any of these interactions will be harmful. Without preventive measures and fluid conditioning, their negative effects can
escalate to the point of component failure. One of the most common failure modes is excessive wear due to loss of
lubrication.

Lubrication and wear


The pressures required in modern hydraulic systems demand sturdy, precisely matched components. And precision
machining leaves very small clearances between moving parts. For example, it is not uncommon for control valves to
have pistons and bores matched and fitted within a mechanical tolerance of ±0.0002 in. (two ten-thousandths of an
inch). In metric units, this is about 5 µm (five millionths of a meter). In modern electrohydraulic devices, tolerances may
be even tighter and clearances can be less than 1 mm. The surface finishes on high-pressure bearings and gears can
result in rolling clearances as small as 0.1 µin.
The hydraulic fluid is expected to create a lubricating film to keep these precision parts separated. Ideally, the film is
thick enough to completely fill the clearance between moving parts, Figure 2. This condition is known as hydrodynamic
or full-film lubrication and results in low wear rates. When the wear rate is kept low enough, a component is likely to
reach its intended service-life expectancy - which may be millions of pressurization cycles.
The actual thickness of a lubricating film depends on fluid viscosity, applied load, and the relative speed of the two
dynamic surfaces. In many applications, mechanical loads are so high that they squeeze the lubricant into a very thin
film, less than 1-µin. thick. This is elastohydrodynamic (EHD) or thin-film lubrication. If loads become high enough, the
film will be punctured by the asperities of the two moving parts. The result is boundary lubrication.
Component and system designers try to avoid boundary lubrication by making sure that fluid has the proper viscosity.
However, viscosity changes as the fluid temperature changes. Also, loads and speed may vary widely during normal
operating cycles. Therefore, most hydraulic components operate at least part of the time with only boundary lubrication.
When that happens, parts of moving surfaces contact each other and are torn away from the parent material. The
resulting particles then enter the fluid stream and travel throughout the system. If not removed by filtration, they react
with other metal parts to create even more wear.
Fluid chemists continually try to minimize potential lubrication problems by improving fluids with additives. Viscosity-
index (VI) improvers are added to help keep viscosity stable as temperature changes. Antiwear additives increase film
strength. If very heavy loads will be applied, the fluid should contain an extreme pressure (EP) additive that reacts with
metal surfaces to form a hard protective film. For fluids in circulating systems, defoamant, demulsifier, detergent, or
dispersant may be added. Rust and oxidation (RO) inhibitors are used in most hydraulic fluids because air and water
are always present to some extent.

Particle-generated wear
The symptoms of wear are diminished system performance and shorter component life. In pumps, wear may first be
detected as reduced flow rate. This is because abrasive wear has increased internal clearance dimensions. Sometimes
called increased slippage, this condition means that the pump is less efficient than it was when new. When pump flow
rate decreases, the fluid system may become sluggish, as evidenced by hydraulic actuators moving slower. Pressure at
some locations in the system also may decrease. Eventually, there can be a sudden catastrophic failure of the pump. In
extreme cases, this can occur within a few minutes after initial start-up of the system.
In valves, wear increases internal leakage. The effect this leakage has on the system depends on the type of valve. For
example, in valves used to control flow, increased leakage usually means increased flow. In valves designed to control
pressure, increased leakage may reduce the circuit pressure set by the valve. Silting interference, Figure 3, causes
valves and variable-flow pump parts to become sticky and operate erratically. Erratic operation shows up as flow and
pressure surges, causing jerky motion in actuators.

Water contamination of liquids


Water in oil-base fluids can be just as destructive as particle contaminants. And water exclusion is very difficult to
accomplish. Because of its affinity for other liquids, water is present in some concentration in most hydraulic systems.
The hygroscopic nature of liquids causes them to pick up a certain amount of water simply from contact with humid air.
When condensation occurs in a reservoir, with subsequent mixing into the base liquid, more water can be added to the
system. Water can even enter with new oil. An oil barrel stored outdoors in a vertical position is likely to have rainwater
collect around its bung. With changes in ambient temperature, some of this moisture can be sucked into the barrel.
Eventually, this water enters the system when the reservoir is filled from that barrel.
Besides these natural phenomena, there are several system- and maintenance-related sources of water. In machine
tool applications, a good deal of water-base coolant can enter hydraulic systems through breather caps and imperfect
seals. Worn and damaged heat exchangers can allow cooling water to leak through seals and ruptured lines into the oil
system - and vice versa.
Each fluid has its own saturation level for water. Below the saturation level, water will be completely dissolved in the
other fluid. For oil-base hydraulic fluids, the saturation level is likely to be in the range of 100 to 1,000 parts per million
(0.01% to 0.1%) at room temperature. At higher temperatures, the saturation level is greater.
Above the saturation level, water becomes entrained, meaning that it takes the form of relatively large droplets. This
also is called free water. Sometimes these droplets combine and precipitate to the bottom of the reservoir. At other
times, due to churning or other mixing action, undissolved water is emulsified so that it exists as very fine droplets
suspended in the oil.

Water's mechanical effects


When water concentration in hydraulic oil reaches 1% or 2%, the response of a hydraulic system may be affected. If
water alters the hydraulic fluid's viscosity, the operating characteristics of the hydraulic system change. When the rate of
water influx is swift, poor system response could be the first indication that water is present.
Cavitation is another symptom of water in the fluid. Because the vapor pressure of water is higher than hydrocarbon
liquids, even small amounts of water in solution can cause cavitation in pumps and other components. This occurs
when water vaporizes in the low pressure areas of components, such as the suction side of a pump. Vaporization is
followed by the subsequent violent collapse of vapor bubbles against metal surfaces in these areas. The loud
characteristic sounds of cavitation may be noticeable when this happens. The result is cavitation damage on the interior
surfaces of hydraulic components because the metal has fatigued.

Emulsified water in oil


Tiny water droplets may be emulsified or suspended in oil-base fluids. Evidence of this is when the fluid appears cloudy
or milky. Sometimes an oil/water emulsion is so tight that it is very difficult to separate the two fluids - even with the
addition of a coalescing agent formulated to do this. While this is desirable in emulsion-type hydraulic fluids, it is highly
undesirable for ordinary oil-base fluids. Some fluid additives encourage emulsification, while others (demulsifiers)
discourage it. The viscosity of a water emulsion can be much different from its original base liquid. As noted earlier,
lower viscosity reduces the thickness of lubricating films, leading to increased wear of surfaces in dynamic contact.
If free water is present in hydraulic fluid and the system operates at temperatures below 32° F, ice crystals may form.
These crystals can plug component orifices and clearance spaces. In hydraulic systems, this will cause slow or erratic
response.
Without fluid analysis to warn of water's presence and appropriate control measures, water content probably will
increase to the point where these and other symptoms appear. Other symptoms include evidence of chemical-reaction
products and eventually, component failures.

Chemical reactions due to water


Water reacts with almost everything in a hydraulic system. Water promotes corrosion through galvanic action by acting
as an electrolyte to conduct electricity between dissimilar materials. The most obvious sign is rust and other oxidation
that appears on metal surfaces. The inside top surface of the reservoir often is the first place rust becomes visible, but
such rust still may go undetected unless the reservoir is drained and opened for servicing. Also, the time it takes for rust
to form depends partly on the original surface treatment used to protect the metal used to build the reservoir.
Unfortunately, before rust is noticed in the reservoir, water probably has damaged other system components. An
inspection of failed bearings and other components may point to corrosion damage. Corroded aluminum and zinc alloys
could have a whitish oxide film. Steel bearing and gear surfaces may show signs of rust and pitting.
Water's reaction with oxidation inhibitors produces acids and precipitates. These water-reaction products also increase
wear and interference. At high operating temperature (above 140° F), water reacts with and actually can destroy zinc-
type antiwear additives. For example, zinc dithiophosphate (ZDTP) is a popular boundary lubricant added to hydraulic
fluid to reduce wear in high-pressure pumps, gears, and bearings. When this type of additive is depleted by its reaction
with water, abrasive wear will accelerate rapidly. The depletion shows up as premature component failure, resulting
from metal fatigue and other wear mechanisms.
Water frequently can act as an adhesive that causes smaller contaminant particles to clump together in a larger mass.
These gluey masses may slow down a valve spool, or cause it to stick in one position. Or these clumps could plug
component orifices. In any case, the result is erratic operation or a complete system failure.

Microbial growth
Over time, water contamination can lead to the growth of microbes - minute life forms such bacteria, algae, yeasts, and
fungi - in the hydraulic system. And the presence of air exacerbates the problem. Microbes range in size from
approximately 0.2 to 2.0 µm for single cells, and up to 200 µm for cell colonies. Left unchecked, microbes can destroy
hydraulic systems just as they destroy living organisms. Under favorable conditions, bacteria can reproduce (doubling
themselves) as rapidly as every 20 minutes. Such exponential growth can form an interwoven mat-like structure that
requires significant shear force to break up. This resistance quickly renders a fluid system inoperable. Besides their
mass volume, bacteria produce acids and other waste products which attack most metals. When this happens, fluid
system performance is degraded, and components fail more rapidly.

Evidence of microbial contamination


The first indication of microbial contamination may be the foul odor that comes from waste and decomposition products
of the microbes. Fluid viscosity may increase due to the mass of material produced by these organisms. At the same
time, the fluid may have a brown mayonnaise-like appearance, or slimy green look.
Unfortunately, by the time these symptoms appear, system components and the fluid itself may be severely damaged.
This could require a major overhaul or replacement of the system.
Properly selected filters will remove microbes. But without adding biocides (substances capable of destroying these
living organisms) to the fluid, fast-growing microbes can place a heavy load on system filters. Combined with wear
debris and chemical reaction products, microbial contamination can result in rapid plugging of filter elements, requiring
frequent replacement.
Water and air are essential for microbe growth. Eliminating water and air from a fluid minimizes microbial problems. But
some systems use water as the base fluid, and air is very difficult to exclude from fluids in operating hydraulic systems.
With water and air present, microbes can usually find some fluid component to feed their growth. When water can not
be controlled by exclusion or removal, biocides should be added to the fluid. A biocide combined with properly selected
water-absorption filters can help minimize chemical-reaction byproducts and microbial contamination.

Exclusion practices
The first defense against fluid contaminants is preventing their entry into a hydraulic system. After that, removing
contaminants before system start-up prevents much damage that can occur early in a system's life. Thereafter, well-
planned routine maintenance will maintain the fluid in peak condition. Here are some of the initial positive steps that can
be taken:
• fit the reservoir with baffles and return-line diffusers, Figure 5, to prevent churning that whips air into the fluid
• equip the reservoir with a breather having an air-filter element with a rating of at least 99% efficiency at 2 µm
• make sure all fittings are properly tightened (besides causing leakage, loose fittings can allow airborne dust to
be sucked into the system)
• flush the system thoroughly before it goes into service
• prefilter fluid before filling the reservoir (it should be as clean as your specification for the system fluid)
• inspect filter indicators to make sure they are working
• use boots and bellows to protect cylinder rods and seals
• replace filter elements before the filter bypass valve opens; otherwise, the system will operate with no filtration
• replace any worn seals and hoses promptly; if not done, the negative effects are the same as loose fittings
• practice good housekeeping whenever a system is opened for maintenance; protect replacement components
from contamination, and
• analyze fluid regularly to detect problems such as overheating, leaking water, clogged heat exchangers,
additive breakdown, etc.

Removal mechanisms
Once contamination is in the fluid, it may be reduced and controlled by settling, outgassing (e.g. in aerated liquids),
filtration/separation, and fluid replacement. The first two mechanisms - settling and outgassing - occur naturally, but
their effect can be enhanced by controlling the fluid environment through system design. The latter two also require
human involvement, again during system design or in maintenance activities after installation.
For settling to occur, a contaminant must have a density greater than the fluid transporting it. The lower the density of a
contaminant particle, the more buoyant it will be in the fluid. The flow rate of the fluid also helps determine how quickly a
contaminant will settle. A contaminant transported by a fluid will stay in suspension if the flow velocity supplies enough
lifting force to overcome gravity. If flow is turbulent, it is more likely that contaminants will stay in suspension.
As mentioned previously, the reservoir can be designed with baffles and return-line diffusers to reduce fluid velocity
enough so that larger particles will settle. On the other hand, contaminants must remain in suspension if they are to be
transported to a filter for removal. This is particularly important in fluid lines and components, where particle settling can
cause unpredictable contaminant removal rates, or silting interference between moving parts. Therefore, system
designers want a reasonable degree of turbulence in the hydraulic system so that smaller particles remain in
suspension. This is as true for the reservoir as elsewhere in the system. A tapered reservoir bottom will help prevent the
collection of smaller contaminant particles due to its reduced bottom surface area and tendency to extend the
turbulence effect. As in many design projects, reservoir construction and piping configuration involves compromises.
Outgassing can be thought of as the inverse of settling. If fluid turbulence is low enough to prevent mixing action,
dissolved air can come out of suspension and rise to the surface of a liquid. Whether the air actually leaves the liquid or
not depends on the relative surface tensions and partial pressures of the air and the liquid. The lower the turbulence in
the reservoir, the more likely it is that a contaminant will leave the fluid by way of outgassing or settling.
Natural mechanisms, such as settling and outgassing, cannot by themselves reduce contamination to an acceptable
level. In the absence of filtration and separation devices, the only alternative is to replace the fluid at periodic intervals.
Even with adequate filtration, fluid replacement cannot be postponed forever. This certainly is true for automotive
lubricants, and points out a fundamental fact of fluid life. There is an economic trade-off between the cost of buying,
installing, and servicing filters and separators, and the cost of replacing the hydraulic fluid more often.

Fluid conditioning objectives


The objective of hydraulic fluid conditioning is to lower total operating costs. If the system can meet or exceed minimum
standards for fluid cleanliness, one or more of these intermediate goals can be achieved:
• reduce maintenance requirements for the fluid system and components
• improve the performance of the system and its fluid
• assure the quality of the final product by improving machine operation, and
• enhance safety and/or reduce risk of injury to personnel (for example, by eliminating the need for
maintenance on or around operating equipment).
Appropriate fluid conditioning increases the mean time between hydraulic component failures. Still, this benefit has to be
properly balanced against the cost of purchasing the filters, replacing elements, and maintaining filtration equipment.
Careful filtration system design and component selection will help minimize these costs. The best way to optimize the
benefit/cost trade-off is to follow sound practices for the selection of filters, elements, and filter media. One general
process is illustrated in the filter-specification flow chart, Figure 5.
Many questions should be answered regarding contaminant removal:
• how clean does the fluid be have to?
• what size particles must be removed?
• how many particles within a given size range need to be removed?
• how efficient must the filter media be in terms of the percentage removal of a given size range - and in terms
of dirt-holding capacity?
• will the fluid contamination stabilize at an acceptable level for a given combination of filters and media?

Component sensitivity
As the flow chart implies, specifiers need to have a feel for the sensitivity of hydraulic components to contaminants of
various sizes and concentrations. Designers and users have observed that some components are more sensitive to
contaminants than others. For example, they may have seen a certain pump quickly fail, while another type lasts for
months in the same system. They also probably have noticed that higher pressures and flow rates tend to make all
components wear out more quickly. Those who are particularly observant may have noticed that the higher the
concentration of airborne contaminants around systems, the sooner they fail. These factors combine to influence the
service life of components.
Another point is that filter media with small pore sizes frequently are more costly, and must be replaced more often than
coarser media. For practical economic reasons, designers must find a compromise between costly ultrafine filtration and
the cost of early component failures. This compromise is to have fluid only as clean as it needs to be, not as clean as
possible.
Designers tend to rely on their own experience as well as information from component manufacturers to determine how
clean hydraulic fluid needs to be. Some conservative manufacturers assume that worst-case conditions exist and
specify a very low acceptable level of contamination for their components. Others take a middle-of-the-road approach,
and specify cleanliness for more or less average conditions.

Additional information sources


Manufacturers' recommendations can be augmented by information that is available from other sources. For example,
OEMs and research laboratories have carried out projects to analyze the sensitivity of pumps, valves, and other
components to contaminants. As a result, guidelines and standards for hydraulic-fluid cleanliness have been published.
These guidelines attempt to interrelate diverse factors such as:
• fluid lubricity (e.g., water-base fluids have lower lubricity than oil)
• abrasiveness of the contaminants commonly found in hydraulic systems
• system duty cycle and cycle rate (high pressure and high cycle rates, combined with contaminants, lead to
earlier fatigue failures)
• component replacement cost
• design life objective in terms of mean time before failure (MTBF); a common goal today is 10,000 hours or
more, and
• degree of risk associated with contaminant-related failures (high risk of personal injury or high cost of lost
production dictates a need for cleaner fluid.)
Fluid variables and system variables both have an effect on a component's sensitivity to contamination. This sensitivity
eventually is reflected in system performance, Figure 6.
The International Standards Organization (ISO) recommends cleanliness levels for various types of components, see
the table below. The levels are stated in terms of industry standards that have been recognized for the past 20 years.
Many fluid power designers apply these recommendations as rules of thumb. Many specifiers now accept and use ISO
4406 (see table on page A/99) as a means of designating the fluid cleanliness required for their systems.

Importance of records
Still, component manufacturers' and industry guidelines should be modified by experience. That requires gathering
enough operating and maintenance data over sufficient time and from enough systems to provide confidence to make
decisions. The data gathered should include the results of regular fluid analysis on systems. The categories of data
collected might include:
Fluid variables - flow, pressure, temperature, and viscosity for circuit branches with the most sensitive or expensive
components.
Fluid analysis - particle counts in various size ranges (e.g., >2, >5, >15, >25, >50, and >100 µm), spectrochemical
analysis (e.g., most likely metals and other contaminants), and water content (% by volume).
Filtration information - model number and manufacturer for the filter(s) and element(s) protecting the circuit for which
other data was gathered; element performance ratings in terms of beta ratios and dirt-holding capacity.
Maintenance data - date system placed in service; dates and descriptions of routine maintenance performed (including
element replacements); reading of the filter element condition indicator (e.g., "needs replacement" or "in bypass"); dates
and descriptions of component failures, including manufacturers' names and model numbers; failure mode analysis
(e.g., fracture, corrosion, wear, etc.) also would be very helpful in determining if contamination was a factor in any
failures.
PC data base and statistical analysis programs also can be used to correlate failures with fluid contamination levels.
This will create a picture of the contamination tolerance of the most sensitive components. It also allows for the
calculation of MTBFs for specific components, certain circuit branches, or the system as a whole.
Obviously, this is data the user must collect. Still, manufacturers can monitor warranty claims as an opportunity to
capture some of this data, and create a clearer picture of component sensitivity. That may cover only the first year or
two of service. A close relationship with customers and distributors can provide an opportunity to gather similar data
over longer periods of time as replacement parts are ordered.

Contamination - dynamic, not static


Another reason for regular fluid analysis is that the contamination level changes with time, and varies by location in the
system. At any point, the amount of contamination present in the fluid depends on three factors:
1. How contamination much was in the fluid when the system was started
2. How much was added to the fluid from all sources during operation (Ingression rate is the term used to describe the
amount of contaminant entering the fluid per unit of time.)
3. How much contamination left the fluid due to all removal mechanisms (settling, and filtration or separation)
These three factors account for the total mass of contaminant in a system at any time. That mass can be calculated
using a material-balance equation:
CT = Ct + Ca - Cs
where:
C is contaminant
T is any point in time
t is time since start of process
Ca is amount added since t
Cs is amount removed since t
The term material balance is used because the equation calculates the net difference between the amount of material or
contaminant entering and leaving the fluid, and adds this difference to what was already there. The calculation applies
to a specific location in the system.
In a circulating system, contaminants not removed will appear at the filter inlet again, along with new contaminant added
to the fluid. This is called a multipass system because the fluid and contaminant make multiple passes through the filter.
As a result, the contaminant concentration in the system fluctuates continuously.
If we consider the initial start-up of a system, the contaminants already present are there as a result of manufacturing
processes or have entered with new fluid. (Each milliliter of fluid out of the original barrel typically contains at least 2,500
particles that are 5 mm and larger in diameter.) A few minutes after start-up, the particulate level will be considerably
higher due to flushing action of the fluid as it flows through new components and piping to pick up debris. Eventually,
more particles enter the system through the reservoir breather and imperfect seals. Still more will be added over time
due to internal wear.

Fluid cleanliness required for typical hydraulic components


Fluid classification
Component type
ISO code
Servovalves 14/11
Vane and piston pumps/motors 16/13
Directional and pressure control valves 16/13
Gear pumps/motors 17/14
Flow control valves and cylinders 18/15
Aircraft test stands 13/10
Injection molding 16/13
Metal working 17/14 - 16/13
Mobile equipment 18/15 - 16/13
New unused oil 18/15
Estimating ingression rate
Assume that a new hydraulic system has been flushed properly before being put
into service. If the system has a multipass filtration system with a given flow rate,
the eventual stabilized level of contaminants will depend on the system's
ingression rate and filter media removal efficiency. If filter efficiency is too low,
the contaminant level will continue to increase due to the wear particles
generated within the system and new particles entering from outside the system.
(This is the scenario in most automobile lubrication systems, and why motor oil
should be changed periodically.) If filter efficiency is high enough, the
contaminant level will decrease and become stabilized, extending the service life
of the hydraulic fluid. Because operating conditions vary, this is a kind of Fig. 7. Simplified representation
dynamic stability. The contaminant level varies within a range determined by of component arrangement for
Multipass Test.
these conditions.
Therefore, to select the appropriate filter media, it is necessary to have some
idea of the ingression rate. Of course the ingression rate probably varies at
different locations in the system, and depends on these factors:
• concentration of ambient airborne contaminants (which enter through worn filler/breathers, loose fittings,
leaking seals, etc.)
• use or absence of an air-filter element in the reservoir breather
• number of components in the system or circuit branch
• types of components that make up the system, particularly if there are rotating components such as pumps
and motors (some types wear faster than others)
• fluid velocity (because higher velocity often may accelerate wear - after flushing is completed)
• system pressure (because higher pressure also tends to increase wear rates)
• fluid temperature (excessive heat can cause fluid and additives to break down, creating contamination), and
• the filter media used (more-efficient media results in lower contaminant levels and reduced wear rates).
This parade of factors makes accurate estimation of ingression rates difficult. An estimate can be made by conducting
particle counts on fluid samples taken from a system with known operating conditions and filtration efficiency. (In
multibranch circulating systems, the reservoir frequently is picked as a convenient location from which to take samples.)
Then, by using a simple filtration model based on the material-balance equation, an ingression rate can be inferred.
Filtration specifiers can construct their own estimates using the same technique. Such an estimate may be more
accurate than one of the averages published in a table. Computerized filtration models also are available that allow a
large number of variables to be quickly manipulated in a kind of ingression-rate What if?analysis.

Cleanliness reference
In order to detect or correct problems, a contamination reference scale is used. Particle counting is the most common
method to derive cleanliness level standards. Very sensitive optical instruments count the number of particles in various
size ranges in a fluid sample. These counts are reported as the number of particles greater than a certain size found in
a specified volume.
The ISO 4406 cleanliness level standard has gained wide acceptance in most industries today. A modified version of
this standard references the number of particles greater than 2, 5, and 15 micrometers in a known volume - usually 1
milliliter or 100 milliliters. (The number of smaller-size particles helps predict silting problems. A high number of larger
particles might indicate catastrophic component failure.)

Filter media
The filter media is that part of the element which actually contacts contaminant and captures it for subsequent removal.
The nature of the particular filter media and the contaminant-loading process designed into the element explains why
some elements last longer in service than others.
During manufacture, media usually starts out in sheet form, then is pleated to expose more surface area to the fluid
flow. This reduces pressure differential across the element while increasing dirt-holding capacity. In some designs, the
filter media may have multiple layers and mesh backing to achieve certain performance criteria. After being pleated and
cut to the proper length, the two ends are fastened together using a special clip, adhesive, or other seaming
arrangement to form a cylinder. The most common media include wire mesh, cellulose, and fiberglass composites, or
other synthetic materials. Filter media is generally classified as either surface- or depth-type.

Surface media
For surface-type filter media, the fluid stream basically flows in a straight path through the element. Contaminant is
captured on the surface of the element which faces the fluid flow. Surface-type elements are generally made from
woven-wire cloth. Because the process used to manufacture the wire cloth can be controlled very accurately, and the
wire is relatively stiff, surface-type media have a consistent pore size. This consistent pore size is the diameter of the
largest hard spherical particle that will pass through the media under specified test conditions. However, during use, the
build-up of contaminant on the element surface will reduce the pore size and allow the media to capture particles
smaller than the original pore-size rating. Conversely, particles (such as fiber strands) that have smaller diameters but
greater length than the pore size may pass downstream through surface media.

Depth media
For depth-type filter media, fluid is forced to take convoluted indirect paths through the element. Because of its
construction, depth-type media has many pores of various sizes formed by the media fibers. This maze of multi-sized
openings throughout the material traps contaminant particles. Depending on the distribution of pore sizes, the media
can have a very high capture rate for very small particle sizes.
The two basic media that are used for depth-type filter elements are cellulose (or paper) and fiberglass. The pores in
cellulose media tend to have a broad range of sizes and are very irregular in shape due to the irregular size and shape
of the fibers. In contrast, fiberglass media consist of man-made fibers that are very uniform in size and shape. These
fibers are generally thinner than cellulose fibers, with a consistently circular cross-section. The differences between
these typical fibers account for the performance advantage of fiberglass media. Thinner fibers can provide more pores
in a given area. Furthermore, thinner fibers can be arranged closer together to produce smaller pores for finer filtration.
Dirt-holding capacity, as well as filtration efficiency, are improved as a result.

Particle counting
Knowing the cleanliness level of the hydraulic fluid in a system is the basis for selecting contamination-control
measures. Particle counting is the most common method of deriving cleanliness-level standards. Very sensitive optical
instruments count the number of particles in various size ranges in a measured fluid sample. These counts are reported
as the number of particles greater than a certain size found in a specified volume of fluid.
The ISO 4406 Cleanliness-Level Standard is accepted in most industries today. A widely used, modified version of this
standard references the number of particles greater than 2, 5, and 15 µm in a known volume - usually 1 or 100
milliliters. The number of particles greater than 2 and 5 µm is a reference point for silt particles, those which can cause
clogging problems. The 15-µm size range indicates the quantity of larger particles present, those which contribute
greatly to possible catastrophic component failure.
To identify a cleanliness level, the number of particles in the sample for each of the three measured sizes is referred to
the ISO 4406 chart, and given an appropriate range number. If a fluid sample contained between 1,300 and 2,500 2-µm
and larger particles (range 18); between 320 and 640 5-µm and larger particles (range 16); and between 40 and 80 15-
µm and larger particles (range 13); the sample would be classified as 18/16/13. Note that the numbers that make up the
ISO cleanliness-code classification will almost never increase as the particle size increases.
Most manufacturers of hydraulic (and load-bearing) equipment conduct tests and then specify an optimum or target
cleanliness level for their components. Exposing components to hydraulic fluid with higher than optimum contamination
levels may shorten the component's service life. It always is best to consult with component manufacturers and obtain
their written fluid-cleanliness-level recommendations. This information is needed in order to select the proper level of
filtration. It also may prove useful for any subsequent warranty claims, as it may draw the line between normal operation
and excessive or abusive operation.

The Multipass Test


The filtration industry uses the ISO 4572 Multipass Test Procedure (also recognized by ANSI and NFPA) to evaluate
filter element performance. During the Multipass Test, Figure 7, fluid circulates through the test circuit under precisely
controlled and monitored conditions. The differential pressure across the element being tested is continuously recorded,
while a constant amount of contaminant is injected upstream of the element. On-line laser particle sensors measure the
contaminant levels upstream and downstream from the test element. (Note that the results of this test are very
dependent on flow rate, type of contaminant, and terminal pressure differential.)
The Multipass Test determines three important element-performance characteristics:
1. Dirt-holding capacity.
2. Pressure differential of the test filter element.
3. Separation or filtration efficiency, expressed as a beta ratio.

Beta ratio
The beta ratio (also known as the filtration ratio) is a measure of a filter element's particle-capture efficiency. Therefore,
it is a performance rating. Here is how a Beta Ratio is derived from Multipass Test results. Assume that 50,000
particles, l0 µm and larger in size, were counted upstream from the test filter, and 10,000 particles in that same size
range were counted downstream from the filter. Substituting in the equation:
βx = (NU)/(ND)
where: x is a specific particle size,
NU is the number of particles upstream, and
ND is the number of particles downstream.
Therefore, β10 = 50,000/10,000 = 5
This result would be read as "Beta ten equal to five." Now, a beta ratio number alone means very little. It is a preliminary
step to finding a filter's particle-capture efficiency. This efficiency, expressed as a percent, can be found by a simple
equation:
Efficiencyx = 100 (1 - 1/β)
Efficiency10= 100 (1 - 1/5) = 80%
In this example, the particular filter element tested was 80% efficient at removing 10-µm and larger particles. For every
five particles in this size range introduced to the filter, four were trapped in the filter media.

Contaminant loading
The term contaminant loading in a filter element refers to the process of filling and blocking the pores throughout the
element. As contaminant particles load the element's pores, fewer open paths remain for fluid flow, and the pressure
required to maintain flow through the media increases. Initially, the differential pressure across the element increases
slowly because plenty of open pores remain for fluid to pass through. The gradual pore-blocking process has little effect
on overall pressure loss.
Eventually, however, successive blocking of media pores significantly reduces the number of pores open for flow. The
differential pressure across the element then rises exponentially as the element nears its maximum life. As the element
continues to load with contaminant, the pressure differential across the filter will continue to increase. This goes on until
the bypass valve (if installed) opens, the element (if without bypass protection) fails structurally, or the clogged element
is replaced. The quantity, size, shape, and arrangement of the pores throughout the element accounts for why some
elements last longer than others.
Consider two of the most common filter media: cellulose and fiberglass, For a given media thickness and filtration rating,
there are fewer pores in cellulose media than fiberglass. Accordingly, the contaminant-loading process would block the
pores of the cellulose media element more quickly than an identical fiberglass media element.
Multi-layer fiberglass media elements are relatively unaffected by contaminant loading for a longer time. The upstream
media has relatively larger pores to capture larger particles; the downstream media layer with very small pores captures
the greater quantity of small particles present in the fluid.

ISO 4406 range numbers


Range Number of particles per milliliter
number More than Up to and including
24 80,000.00 160,000.00
23 40,000.00 80,000.00
22 20,000.00 40,000.00
21 10,000.00 20,000.00
20 5,000.00 10,000.00
19 2,500.00 5,000.00
18 1,300.00 2,500.00
17 640.00 1,300.00
16 320.00 640.00
15 160.00 320.00
44 80.00 160.00
13 40.00 80.00
12 20.00 40.00
11 10.00 20.00
10 5.00 10.00
9 2.50 5.0
8 1.30 2.50
7 0.64 1.30
6 0.32 0.64

General comparison of filter media


Dirt-
Capture Differential Service Initial
Media material holding
efficiency pressure life cost
capacity
Fiberglass High High Moderate High Moderate to high
Cellulose (paper) Moderate Moderate High Moderate Low
Low Low Low Low Moderate Moderate to high
Cleanliness required for typical hydraulic components
Component ISO code
Servovalves 16/14/11
Proportional valves 17/15/12
Vane and piston pumps and motors 18/16/13
Directional and pressure control valves 18/16/13
Flow control valves and cylinders 20/18/15
New unused fluid 20/18/15

Beta ratios and


capture efficiencies
Beta ratio Capture efficiency
(at a given (at the same
particle size) particle size)
1.01 1.0%
1.10 9.0%
1.50 33.3%
2.00 50.0%
5.00 80.0%
10.00 90.0%
20.00 95.0%
75.00 98.7%
100.00 99.0%
200.00 99.5%
1,000.00

Filter-element life profile


Every filter element has a characteristic relationship between pressure differential and contaminant loading. This
relationship can be defined as the filter element life profile. The actual life profile obviously is affected by the system
operating conditions. Variations in the system flow rate and fluid viscosity affect the clean pressure differential across
the filter element and have a well-defined effect upon the actual element life profile.
The filter element life profile is very difficult to evaluate in actual operating systems. The system's ratio of operating time
to idle time, the duty cycle, and the changing ambient contaminant conditions all affect the life profile of the filter
element. In addition, precise instrumentation for recording the change in the pressure loss across the filter element
seldom is available. Most machinery users and designers simply specify filter housings with differential-pressure
indicators to signal when the filter element should be changed.
Multipass Test data can be used to develop the pressure-differential-versus-contaminant-loading relationship. As
mentioned, operating conditions such as flow rate and fluid viscosity affect the life profile of a filter element. Life profile
comparisons can be made only when all these operating conditions are identical, and the filter elements are the same
size.
Under those conditions, the quantity, size, shape, and arrangement of the pores in the filter element determine the
characteristic life profile. Filter elements manufactured from cellulose media, single-layer fiberglass media, and multi-
layer fiberglass media have very different life profiles.

Filter housings
The filter housing is the pressure vessel which contains the filter element. It usually consists of two or more
subassemblies, such as a cover (or head) and a removable bowl that allows access to the element. The housing has
inlet and outlet ports that enable fluid to enter and leave. Housing options may include bypass valves and element-
condition indicators.
Primary concerns in the housing-selection process include mounting methods, porting options, indicator options, and
pressure rating. Except for the pressure rating, all depend on the physical system arrangement and the preferences of
the system designer. Pressure rating of the housing is far less arbitrary; it is determined by system needs before the
housing style is selected.

Pressure ratings
Location of the filter in the circuit is the primary determinant of pressure rating of the component. Filter housings are
generically designed for one of three locations in a circuit: suction, pressure, or return lines. One characteristic of these
locations is their maximum operating pressures. Suction and return line filters are generally designed for lower
pressures - 500 psi or less. Pressure filter locations may require ratings from 1,500 to 6,000 psi.
Note that it is essential to analyze the circuit for pressure-spike potential as well as steady-state conditions. Some
housings have restrictive or lower fatigue pressure ratings. In circuits with frequent high-pressure spikes, another type
housing may be necessary to prevent fatigue-related failures.

Bypass valves
Bypass valves open flow paths around filter elements to prevent their collapse or bursting when they become heavily
loaded with contaminant. As contaminant builds up in the element, the differential pressure across the element
increases. At a pressure well below the failure point of the filter element, the bypass valve opens, allowing flow to go
around the element. Some bypass valve designs have a bypass-to-tank option. This directs the unfiltered bypass flow
back to the tank through a third port, preventing unfiltered bypass fluid from entering the system. Bypass valves also
prevent pump cavitation when used with suction line filters. When specifying a bypass-type filter, it generally can be
assumed that the manufacturer has designed the element to withstand the bypass valve differential pressure when the
bypass valve opens.
Note that some of the upstream contaminant particles also bypass the filter element with the fluid and enter the
downstream system. When this happens, the effectiveness of the filter element is compromised and the attainable
system fluid cleanliness degrades.
Other filters are designed specifically with no bypass valve (sometimes called a blocked bypass). They prevent any
unfiltered flow from going downstream, thus protecting servovalves and other contaminant-sensitive components. In
filters without bypass valves, higher collapse-strength elements may be required, especially if installed in high-pressure
locations. When specifying a non-bypass filter design, make sure that the element has a differential-pressure rating
close to the maximum operating pressure of the system, and that the filter has a condition indicator.
After a housing style and pressure rating are selected, the bypass valve setting needs to be chosen. This setting must
be established before sizing the filter housing. Everything else being equal, the highest bypass cracking pressure
available from the manufacturer should be specified. This will provide the longest element life for a given filter size.
Occasionally, a lower setting may be selected to help minimize energy loss in a system, or to reduce backpressure on
another component. In suction filters, either a 2- or 3-psi bypass valve is used to minimize the chance of potential pump
cavitation.

Element-condition indicators
The element-condition indicator signals when the element is loaded to the point that it should be cleaned or replaced.
The indicator usually has calibration marks which also indicate if the bypass valve has opened. The indicator may be
linked mechanically to the bypass valve, or it may be an entirely independent differential-pressure sensing device.
Indicators may give visual or electrical signals or both. Generally, indicators are set to trip at a differential pressure
anywhere from 5% to 25% below that which opens the bypass valve.

Sizing housing and element


The filter housing size should be large enough to achieve at least a 2:1 ratio between the bypass valve setting and the
pressure differential of the filter with a clean element installed. For longer element life, this ratio should be 3:1 or even
higher.
Referring to typical flow/differential-pressure curves from a manufacturer's catalog, the filter specifier needs to know the
operating viscosity of the fluid and the maximum (not the average) flow rate to assure that the filter does not spend a
high portion of time in bypass due to flow surges. This is particularly important in return-line filters, where flow
multiplication from large cylinders may increase the return flow compared to the pump's flow rate.
Always consider ambient temperature conditions when sizing filters. Low ambient temperatures may increase fluid
viscosity to the point where pressure differential across the filter assembly also may increase considerably.
If a filter was fitted with a 50-psi bypass valve, the initial (clean) pressure differential should be no greater than 25 psi
and preferably 16 2/3 psi or less. These pressures are calculated from the 3:1 and 2:1 ratios of the 50-psi bypass
setting and initial pressure differential.
Standard filter assemblies normally are manufactured with a bypass-valve cracking pressure between 25 and 100 psi.
The bypass valve in most of these assemblies actually limits the maximum pressure drop across the filter element. As
the element becomes blocked with contaminant, the pressure differential increases until it reaches the bypass valve
cracking pressure. At this point, part of the flow through the filter assembly begins to bypass the element through the
valve. This action limits the maximum pressure differential across the filter element.
The relationship between the starting clean pressure differential across the filter element and the bypass valve pressure
setting must be considered. A cellulose element has a narrow region of exponential pressure rise. For this reason, the
relationship between the starting clean pressure differential and the bypass valve pressure setting is very important.
This relationship in effect determines the element's useful life.
In contrast, the useful element life of single-layer and multi-layer fiberglass elements is established by the nearly
horizontal, linear region of relatively low pressure drop increase, not the region of exponential pressure rise.
Accordingly, the filter assembly's bypass valve cracking pressure, whether 25 or 75 psi, has relatively little impact on the
useful life of the element. Thus, the initial pressure differential and bypass valve setting is less a sizing factor for
fiberglass media.

Filter types and locations


The type of filter - suction, return, pressure, or off-line - and its physical location in the circuit are almost inseparable by
definition.
Suction filters serve to protect the pump from fluid contamination. They are located upstream from the pump's inlet
port. Some may be simple inlet strainers, submersed in fluid in the tank. Others may be mounted externally. In either
case, suction filters have relatively coarse elements, due to cavitation limitations of pumps. (Some pump manufactures
do not recommend the use of a suction filter. Always consult the pump manufacturer for inlet restrictions.) For this
reason, suction filters are not used as a system's primary protection against contamination, and in fact, the use of
suction strainers and filters has greatly decreased in modern hydraulic equipment.
Return filters may be the best choice if the pump is particularly sensitive to contamination. In most systems, the return
filter is the last component through which fluid passes before entering the reservoir. Therefore, it captures wear debris
from all of the system's working components and any particles that enter through worn cylinder rod seals before such
contaminant can enter the reservoir and be pumped back into the system. Because this filter is located immediately
upstream from the reservoir, its pressure rating and cost can be relatively low.
Note that retracting some cylinders with large diameter rods may result in flow multiplication. This high return-line flow
rate may open the filter bypass valve, allowing unfiltered fluid to pass downstream. This probably is an undesirable
condition and should be considered when specifying the filter.
Pressure filters are located downstream from the system pump. They are designed to handle the system pressure and
are sized for the specific flow rate in the pressure line where they are located. Pressure filters are especially suited for
protecting sensitive components, such as servovalves, directly downstream from the filter. Because pressure filters are
located just downstream from the pump, they also help protect the entire system from any pump-generated
contamination.
Duplex filters, a common special configuration, may include both pressure and return filters. Duplex filters provide
continuous filtration. They have two or more filter chambers and include the necessary valving to allow for uninterrupted
operation. When one filter element needs to be serviced, the duplex valve is shifted, diverting flow to the opposite filter
chamber. The dirty element can then be changed, while flow continues to pass through the cleaner element. The duplex
valve typically is an open cross-over type, which prevents any flow blockage.

Off-line filtration
This increasingly popular filtration arrangement - also referred to as recirculating, kidney loop, or auxiliary filtration - is
totally independent of a machine's main hydraulic system. This makes it attractive as a retrofit project for problem
systems. An off-line filtration circuit includes its own pump and electric motor, a filter, and the appropriate connecting
hardware. These components are installed off-line as a small subsystem separate from the working lines, or they may
be included in a fluid-cooling loop. Fluid is pumped continuously out of the reservoir, through the off-line filter, and back
to the reservoir. A rule of thumb: the off-line pump should be sized to flow a minimum of 10% of the main reservoir
volume.
With its polishing effect, off-line filtration is able to maintain the system's fluid at a constant contamination level. As with
a return line filter, the off-line loop is best suited to maintain overall system cleanliness; it does not provide protection for
specific components. An off-line filtration loop has the added advantage that it is relatively easy to retrofit on an existing
system that has inadequate filtration. Also, the off-line filter can be serviced without shutting down the main system.

Descriptive factors for particles


Particulate contamination may be characterized by the following identification factors.
agglomeration the tendency of particles to bond together. This action is generally detrimental in fluid-contamination
control.
compaction degree of packing from sedimentation process. As void spaces decrease and bulk density increases, silt
condition intensifies.
concentration weight per unit volume of fluid or number of particles greater than given size per unit volume of fluid.
density mass of particle per unit of volume. Density affects the rate at which particles settle out from the fluid.
dispersion the tendency of particles to remain separated. This is a factor in particle separation and analysis.
hardness resistance to abrasion and the particles potential to abrade exposed surfaces.
settling terminal velocity of particles controls the degree of particle suspension provided by the flowing fluid.
shape degree of irregularity of particle structure or topography. A factor important to the cutting or abrading ability of the
particle.
size structural extent of particle as defined by geometric, derived, and hydrodynamic diameters. Such diameters have
significance on a statistical basis.
size distribution frequency of occurrence of each particle size in the population. Cumulative particle size distribution
curves are the most popular type in fluid-contamination control.
size limits the size range in which only fractureless deformation occurs and the lower filtration limit of interest.
state condition where size and shape cannot be altered without forceful shearing of crystalline or molecular bonds. A
concept important to the understanding of particle generation and growth.
transport the life force needed to overcome the buoyant weight of the particle. When this is achieved, the flow conduit
does not retain particles on its surface.

You might also like