You are on page 1of 29

MNL37-EB/Jun.

2003

Hydraulic Fluids
W. A. Givens^ and Paul W. Michael^

T H E PRIMARY PURPOSE OF A HYDRAULIC FLUID is to

Where:
K = Bulk modulus
Vo = Original volume
AP = Pressure change
Ay = Change in volume

transfer

power. The concept of fluid power is based on a principle articulated by Blaise Pascal, which is usually given as follows:
"Pressure applied to an enclosed fluid is transmitted undiminished to every portion of that fluid and the walls of the
containing vessel" [1]. Within the context of fluid power,
pressure is related to the force acting on a confined fluid as
illustrated in Fig. 1 [2]. This principle has given rise to mode m hydraulics, which entails highly engineered systems for
efficiently controlling fluid flow to transfer energy and accomplish work.
The heart of any hydraulic system is the pump, which pulls
in fluid from a reservoir by creating a vacuum at its inlet and
then forces the fluid through its outlet, usually against pressure created by flow controllers and/or actuators downstream of the p u m p . Pumps, actuators, and other system
components have surfaces that move relative to each other,
often at high speeds, pressures, and temperatures. These
components require cooling and lubrication for efficient performance and durability. Consequently, hydraulic fluids not
only must transmit power, they serve critical functions as lubricant and heat transfer medium.

Heat Transfer
Heat is generated as a by-product of normal operation of a
hydraulic circuit. Friction between the moving parts of a
p u m p or hydraulic motor, as well as friction between the
fluid and surfaces of valves, pipes, and other circuit devices
generates heat. In addition, heat is generated in a hydraulic
system as a result of the dissipation of the potential energy of
pressurized fluid [8]. As a hydraulic fluid is circulated
through a system, heat is transferred from high temperature
areas to coolers, reservoirs, and other regions of the circuit
where it is dissipated. As can be seen in Table 1, typical specific heat and thermal conductivity values for hydraulic oils
are a fraction of that of water [4]. These factors are an important consideration in sizing hydraulic system coolers because the inherent cooling efficiency of petroleum based hydraulic fluid is less than that of water. ASTM D 2717, Test
Method for Thermal Conductivity of Liquids and ASTM D
2766, Test Method for Specific Heat of Liquids and Solids are
used to determine these properties of fluids.

P o w e r Transfer
To transfer power efficiently, a hydraulic fluid must exhibit
minimal compressibility. Low compressibility allows all of
the pressure applied to the fluid to be available for direct
and effective transmission to system components such as
motors, cylinders, or other actuators. The compressibility of
a fluid is generally discussed in terms of its "bulk modulus,"
which describes the change in fluid volume as a result of applied pressure [3]. The bulk modulus of a fluid, which is the
reciproccd of compressibility, is described by Eq 1. There
are a n u m b e r of m e t h o d s available for estimating the
isothermal secant bulk modulus of a fluid based upon its
viscosity and density characteristics [4,5]. As depicted in
Fig. 2, the bulk modulus for oil also varies with temperature
[6]. For petroleum oils, compressibility is often assumed to
be 0.5% for each 1000 psi pressure increase u p to 4000 psi
[7].
Bulk modulus {K) = -Vo

(\PI\V)

Lubrication
The durability of hydraulic equipment depends to a large extent upon the lubricating properties of the fluid. As a lubricant, the key function of the hydraulic fluid is to reduce friction between contact surfaces. A reduction in friction lowers
contact t e m p e r a t u r e s a n d wear. This is accomplished
through a combination of hydrodjoiamic and boundary lubrication mechanisms. The hydrodynamic lubricating properties of a fluid are governed by its physical properties while
boundciry lubrication is a function of fluid chemistry. A discussion of hydraulic fluid wear testing is presented in the
Wear Protection section of this chapter.

(1)

TRENDS

* Exxon Mobil Research & Engineering, Paulsboro Technical Center, 600 Billingsport Rd., Paulsboro, NJ 08066.
^ Benz Oil, 2724 West Hampton Avenue, Milwaukee, WI 53209.

353

A brief outline of major trends in the motion control industry, particularly with respect to hydraulic equipment design
and fluid requirements, is presented as a backdrop for the
discussion of hydraulic fluids test methods. As motion con-

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
Copyright' 2003 by A S I M International
www.astm.org
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

354 MANUAL 3 7: FUELS AND LUBRICANTS HANDBOOK

Force

Area

F = force in pounds
p = pressure in pounds / sq. incli (psi)
A = sq. in.

FIG. 1Relationship of force, pressure, and area in fluid power. Any one of the parameters equals the other two in the relationship depicted by the triangle.
TABLE 1Thermal conductivity and specific
heat values for oil and water.

40

30

M
Q.

Oil
Water

z
CO

Thermal
Conductivity
Btu/h/ft^/F/Ft
@ 212F

Thermal
Conductivity
W/m-K
@373K

Specific
Heat
BTU/lbF
@68F

Specific
Heat
J/kg K
@293K

0.08
0.39

0.14
0.67

0.47
1.0

1966
4184

20

I
10
3

m
100

200

300

FIG. 2- -Effect of temperature on the bulk modulus of petroleum


fluid.

trol technology advances, there is a trend towards higher performance and efficiency. For hydraulic equipment, this
translates into a concentration of horsepower in smaller
components. There are a n u m b e r of reasons for such a trend.
Equipment manufacturers are looking for ways to minimize
raw material usage and cost. Users of the equipment demand
smaller systems for better space utilization in industrial environments cind compact multifunctional capabilities in mobile equipment. These advancements in mechanical design
along with e n c r o a c h m e n t of environmental, health, and
safety regulations fuel the following trends:
Hydraulic equipment builders will continue to push comp o n e n t manufacturers to design parts to a c c o m m o d a t e
high pressures a n d t e m p e r a t u r e s . F o r example, hoses,
valves, and other fittings will continue to evolve in terms of
materials used as well as actual functional design.

Smaller c o m p o n e n t s will m e a n smaller p u m p displacements [cubic inches or cc per p u m p revolution]. To maintain flow rates at present or higher levels, p u m p speeds will
be increased [cubic inches/minute = displacement X speed
(rpm)]. Smaller reservoir sizes will mean shorter fluid residence times and will therefore dictate use of hydraulic fluids with improved air release characteristics.
Smaller dimensional clearances will be required. These
smaller clearances will dictate more stringent fluid cleanliness requirements to prevent abrasive wear from particulate c o n t a m i n a n t s a n d failure of servo or proportional
valves.
Fluid cleanliness will increasingly be emphasized as an effective way of increasing equipment durability and controlling warranty costs. As a result, users will move to finer
filtration and specify pre-filtered hydraulic fluids [9]. Consequently, the filterability of the hydraulic fluid will continue to grow in significance. (Filterability is described in
section 4.6.)
Quieter hydraulic systems will be required in order to meet
workplace noise restrictions and compete with electric motors. Reduction of noise levels in hydraulic equipment has
been attained by the insulation that absorbs the noise. This
insulation results in higher system temperatures, as heat is
not as readily dissipated.
Components and actuators, such as cylinders, will be designed with tighter seals to increase efficiency and reduce

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER

13: HYDRAULIC

FLUIDS

355

leakage. The effects of this trend include increased stress


on seal materials and cylinder chatter resulting from reduced lubrication between seal and cylinder wall. In addition, certain applications will require fill-for-life systems
that translate into lower maintenance and disposal costs.
Consequently, fluids will remain in a system for longer periods, since meike-up fluid is not required.
A growing awareness of the environmental impact of
chemicals will lead to further restrictions on performance
additives eind base stocks used in lubricants. As a result,
lubricant p r o d u c e r s are required to address such issues through alternative (usually more costly) chemistry
and the development of environmentally friendly (nontoxic/biodegradable) lubricants.
The hydraulic fluid industry has evolved from the use of
plain water in hydraulic systems to the use of advanced
fluid technologies that continue to evolve as performance
requirements become more stringent and equipment designs become more sophisticated [10]. Due to environmental health and safety issues, hydraulic systems are once
again being designed to employ pure water as hydraulic
fluid [11].

Solvent refining yields base oils that fall into Group I while
hydroisomerization and deep hydrogenation processes yield
low sulfur, high paraffin content Group II a n d Group III
base stocks. Because of their lower aromatic and sulfur content, hydraulic fluids formulated from Group II and Group
III base stocks typically have superior oxidation stability.
However, more highly refined stocks tend to be less effective
at dissolving additives. Not only is additive solubility a concern, additive chemistries and their functional mechanisms
may be b o t h synergistic a n d antagonistic. Thus, additive
chemistry must be ceirefully balanced to achieve optimum
performance. In the following section, test methods for evaluating key fluid properties such as oxidation stability, wear
prevention, and corrosion inhibition are discussed. These
methods have been developed to measure characteristics of
hydraulic fluids that are thought to correlate to performance
in "real-life" applications as well as gage additive response
for the fluid formulator. In order to provide a link between
fluid tests and additive chemistry, a description of the generally accepted functional mechanisms of additives is also
included.

PETROLEUM BASE STOCKS

FLUID CHARACTERISTICS AND


PERFORMANCE

Most hydraulic fluids consist of a base fluid and additives


that are designed to i m p a r t chemical characteristics and
functionality to the finished product. Operating conditions
and equipment builder specifications generally dictate the
type of fluid that is needed and thus, the kind of base stocks
and additives employed. In petroleum based hydraulic fluids
the typical concentration of additives is less than 3.0% by
weight. Paraffinic oils are the primary base stock utilized in
hydraulic fluids but other materials, from polyglycols to vegetable oil, serve as the basis for formulating hydraulic fluids.
From a historical standpoint, solvent reflned paraffinic oils
have been the most widely used base stock for hydraulic applications. In recent years alternative refining processes such
as catalytic isomerization and deep hydrogenation have been
developed to yield higher purity base oils that are better
suited to withstand severe operating conditions [12]. These
base stocks are categorized by the American Petroleum Institute (API) according to their composition and viscosity index
[13]. Groups I through III consist of crude derived base oils
while Group IV is reserved for synthetic polyalphaolefins.
Low viscosity index naphthenic oils and other base stocks
that do not meet Group I through IV criteria are classified
as Group V. The API Base Oil classification is described in
Table 2.
TABLE 2API base oil classifications.
Category
Group I
Group II
Group III
Group rV
Group V

Composition
< 9 0 % Saturates or
> 1 0 % aromatics
9 0 % Saturates or
< 1 0 % aromatics
> 9 0 % Saturates or
< 1 0 % aromatics
All polyalphaolefins (PAO)
All others not included in
Groups 1,11, m or IV

Suli^ir
>0.03%

Viscosity
Index
80-120

<0.03%

80-120

<0.03%

>120

Oxidation a n d Thermal Stability


An important characteristic of a hydraulic fluid is its ability
to withstand high temperatures. This is because horsepower
losses in hydraulic systems directly result in transfer of heat
to the fluid. Resulting high temperatures can cause hydraulic
fluids to react with oxygen. The rate of this reaction accelerates exponentially with increasing temperatures and is further catalyzed by metals like copper and iron, especially at
temperatures above 200F [14]. Rate constants for the oxidation of saturated hydrocarbons at 125C are as much as 40
times higher than rate constants at 60C [15]. Thus, fluid oxidation is highly dependent upon hydraulic system operating
temperatures. Lubricants expected to operate in high temperature environments are tjrpically fortified with additives
known as antioxidants, which are discussed in the Antioxidants section. Oxidative stabilization of the fluid translates
directly into extended oil service life. Failure to resist oxidation can result in thickening of the oil (viscosity increase),
formation of acidic byproducts, and subsequent deposit formation.
Not only can heat cause oxidation, fluids may thermally degrade upon exposure to high temperatures with litde or no
oxygen present. The thermal stability of a hydraulic fluid is
dependent mainly on the intrinsic ability of the base fluid or
its components to resist decomposition at high temperatures.
Unlike oxidation, controlled thermal degradation of certain
types of additives [such as Zinc Dialkyldithiophosphate
(ZDTP)] is desirable, because it is the very mechanism by
which they react with the metal surfaces they are designed to
protect [16]. Similar to oxidation however, the negative effects of thermal degradation may include increased acidity,
thickening of the oil, and deposit formation. Therefore, good
control of thermal degradation results in the retention of desired fluid properties.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

356

MANUAL

37: FUELS AND LUBRICANTS

High Temperature/Oxidation

HANDBOOK

Tests

One of the most commonly sited methods for measuring the


abihty of a fluid to resist oxidation is ASTM D 943, Standard
Test Method for Oxidation Characteristics of Inhibited Mineral Oils (also known as the Turbine Oil Oxidation Stability
Test - TOST). In this test, 300 ml of fluid a n d 60 ml of distilled
water are placed in a large test tube together with coils of
copper and iron wire (Fig. 3). The fluid is heated to 95C
(203 F) and oxygen is bubbled through the fluid at a controlled rate. The test is complete when the Total Acid Number (AN) of the fluid reaches 2.0 mg KOH/g. As can be seen
from the reaction scheme in Fig. 5, cJdehydes eire among the
chemical by-products of hydrocarbon oxidation. These aldehyde compounds are readily converted to carboxylic acids in
the hydraulic system [17,18]. Since carboxylic acids are corrosive to yellow metals and agglomerate to form deposits,
they have a detrimental effect upon fluid performance when
their concentration becomes excessive. The concentration of
acidic oxidation debris present in a fluid can be determined
by titration with potassium hydroxide. For the D 943 test, a
variation on ASTM D 664 Acid N u m b e r of Petroleum Products by Potentiometric Titration is used. This method, ASTM
D 3339, Test Method for Acid N u m b e r By Semi-Micro Color
Indicator Titration is utilized because it permits a 0.2-1.0 g
sample size for total acid numbers in the 0.5-3.0 mg KOH/g

OXYGEN
DELIVERY
TUBE

range. The hours to form 2.0 m g of KOH equivalents of acidic


oxidation products per gram of some typical fluids are shown
in Table 3. In general, turbine oils provide longer TOST oxidation life t h a n antiwear hydraulic fluids because turbine
oils typically do not contain zinc dialkyldithiophosphate
(ZDTP). Zinc dialkyldithiophosphate r e d u c e s the t i m e it
takes for a fluid to reach 2.0 mg KOH/g because it is acidic
and its mere presence raises the acid n u m b e r of the fluid. In
addition ZDTP is subject to hydrolysis and forms acidic compounds as it degrades. Ester based fluids such as rapeseed
oils are also subject to hydrolysis, which accounts for their
poor performance in the D 943 test. When the D 943 test is
run without water (dry method), the oxidation life of a synthetic ester can be extended by nearly a factor of 100.
The a m o u n t of sludge produced in the TOST test may be
measured by ASTM D 4310, Test Method for Determination
of the Sludging Tendencies of Inhibited Mineral Oils. In this
test, the fluid is subjected to D 943 test conditions for 1000 h.
At the end of this time, the sludge produced is determined
gravimetrically by filtration of the oxidation tube contents
through 5-/u,m pore size cellulose acetate filter disks. To a certain extent the D 943 and D 4310 tests evaluate different
mechanisms of high temperature degradation. In the D 943
test, acidity is measured and this acidity is predominantly
due to formation of carboxylic acids by the conventional liquid phase oxidation mechanism shown in Fig. 4. In essence
D 943 measures the stability of the base oils and the effectiveness of oxidation inhibitors. Sludge formation in hydraulic oils is to a greater extent due to theimal degradation
of the ZDTP antiwear additive. Consequently, the result of a
D 4310 test is an indication of the thermal stability of ZDTP.
Figure 5 shows a model for the mechanism of sludge formation by zinc dialkyldithiophosphate [19].
Another method for measuring the sludging tendency of
hydraulic fluids is the Cincinnati Machine Heat Test [20].
This test has been adopted as an ASTM procedure and is designated ASTM D 2070, Standard Test Method for Thermal
Stability of Hydraulic Oils. In this test, polished pre-weighed
copper and steel rods are placed in a beaker containing 200
cc oil and heated to 135C (275F) for 168 h. At the end of the
test, the copper and steel rods are examined for discoloration
due to corrosion caused by carboxylic acids and sulfur compounds formed by thermal degradation. Sludge content and
viscosity increase are also measured (Table 4).
Antioxidants

CATALYST
COILS

Oxidation inhibitors, commonly referred to as antioxidants,


are chemicals that reduce the rate at which oxidative degradation of a lubricant occurs. Degradation begins with the reaction of hydrocarbon molecules at elevated temperatures to
form unstable reactive species known as free radicals. These
TABLE 3D 943 turbine oil oxidation test life of
typical hydraulic fluids.

FIG. 3Oxidation cell and sampling tube for


ASTM D 943 apparatus.

Fluid Type

Hours to TAN of
2.0 by D 943 Method

Synthetic ester without antioxidant


Mineral oil without additives
Antiwear hydraulic oil. Group I base stock
Antiwear hydraulic oil, Group 11 base stock
Synthetic ester with antioxidant, dry method
R & O hydraulic oil

65
300
2016
5040
5500
>10,000

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER 13: HYDRAULIC FLUIDS 357


Temperature
Light, catalyst

Initiation

RH

Propagation

R + O2

ROO*

Peroxy radical

ROO + RH

ROOH + R*

Hydroperoxide

ROOH

-*

RO + OH

Alkoxy radical

RO + RH

->

ROH + R

Alcohol

OH + RH

->

H2O + R

Water

Branching

Termination

Alkyl radical

R + ROO

Alcohols

RO + ROO

Aldehydes

ROO + ROO

Ketones

RO + R

Acids

R + R

Longer chain
hydrocarbons

FIG. 4Reaction scheme for liquid hydrocarbon oxidation.

Hydraulic Oil
RO

RO

SZnS

OR

OR

Base Oil
(Paraffinic)
and
Additives

Machines
and Outside
Environment

T
Reaction withi

Thermal P^''^^
Deterioration
Degradation
witii Water
ZnSq
RO

Decomposition
Oxidation
Reaction with
l\^etai Ions

Polyphosphates
0

OR

0- -Zn0

OR

\ ^
RO

Oxidation
Products
and
Metal Soaps

Wear Particles,
Dust, Rust, Water
andOtliers

Sludge
FIG. 5Mechanism of sludge formation by zinc dialkyldithiophosphate.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

358 MANUAL 37: FUELS AND LUBRICANTS HANDBOOK


TABLE 4Cincinnati machine thermal stabihty
test performance requirements.
Property

Condition of steel rod


Visual
Deposits
Corrosion
Condition of copper rod
Visual
Corrosion
Condition of fluid
Viscosity
Sludge
Total acid number

Requirement

No discoloration
3.5 mg maximum
1.0 mg maximum
5 rating maximum
10.0 mg maximum
5% change maximum
25 mg/100 mL max
50 % maximum

species react with oxygen and non-oxidized oil to form additional free radicals, which propagate the oxidation process.
This generally accepted mechanism is described as free radical chain reaction and is illustrated by the steps shown in
Fig. 4.
Antioxidants interrupt this chain reaction and thus, reduce
the rate of oxidation and the resulting viscosity increase and
acid and deposit formation. There are two general mechanisms by which these additives inhibit oxidation. The antioxidants are therefore categorized as primary or secondary, depending u p o n the m e c h a n i s m of oxidation inhibition.
Primary antioxidants, commonly referred to as "free radical
scavengers," react with the peroxy radicals and hydroperoxides to form inactive compounds (Fig. 6) [21]. Examples of
primary antioxidants include hindered phenols and aromatic
amines. Secondary antioxidants, commonly referred to as
"peroxide decomposers," react with hydroperoxides or peroxy radicals to form less reactive compounds. Examples of
secondary antioxidants include sulfur a n d / o r p h o s p h o r u s
c o m p o u n d s a n d metal dithiophosphates (Fig. 7). Antioxidants genereilly function in the bulk lubricant and are consumed as they do their job [22].

Detergents

IDispersants

Detergents and dispersants are used to delay formation and


subsequent deposit of insoluble oil degradation species. The
terms detergent and dispersant are often used interchangeably, but are generally differentiated by their composition
and primary functionality. Detergents are metallo-organic
compounds that neutralize acidic deposit precursors, while
dispersants are predominantly organic chemicals that keep
insoluble materials dispersed a n d suspended in the lubricant.
The t e r m "ashless" dispersants, m e a n i n g non-metallic, is
used to further differentiate dispersants from detergents.
Some detergents have the ability to disperse and suspend insolubles, while some dispersants are capable of neutralizing
precursors of deposits. Typical lubricant detergents include
barium, calcium, and magnesium phenates, phosphates, salicylates a n d sulfonates. Ashless dispersants are typically
alkyphenol-based or alkyl succinimides.

(R0)3P?-0O3
(R0)3P + R'OOH
H
(R0)3P=0 + HOR'
FIG. 7Secondary antioxidants such as the phosphite compound depicted above inhibit oxidation by decomposing hydroperoxides. This prevents the oxidation process from progressing beyond the branching stage In the reaction
mechanism.

o
+ R00
ROO^^R
FIG. 6Reaction scheme for primary antioxidants. Primary or freeradical trapping antioxidants work by donating a hydrogen radical H* to
the peroxy radical formed during mineral oil oxidation. Due to steric hindrance, the antioxidant radical does not attack mineral oil molecules,
i.e., R-H bonds. Consequently, the radical chain is terminated.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER 13: HYDRAULIC FLUIDS


Wear Protection
Reduction of friction and prevention of wear is the fundamental purpose of a lubricant. Lubricants reduce friction in
machine components by producing a physical or chemical
barrier between surfaces that slide or roll past each other. Depending on equipment design and function, lubricants function within three commonly recognized regimes: hydrodynamic, mixed-film, and boundary lubrication (Fig. 8) [23].
Hydrodynamic lubrication is often the dominant lubrication regime under conditions of moderate temperatures and
loads. According to the ASM Handbook on Friction, Wear
and Lubrication Technology, [24] hydrodynamic lubrication
is "a system of lubrication in which the shape and relative
motion of the sliding surfaces causes the formation of a fluid
film that has sufficient pressure to separate the surfaces." In
this regime, viscosity is the most important fluid characteristic because it, in combination with sliding speed, contact geometry and load, determines the thickness of the lubricating
film, and determines whether or not the surfaces will contact
each other.
Fluid viscosity plays an important role in hydraulic applications. A hydraulic fluid that is too low in viscosity will
cause low volumetric efficiency, fluid overheating, and increased pump wear. A hydraulic fluid that is too high in viscosity will cause poor mechanical efficiency, difficulty in
starting, and wear due to insufficient fluid flow [25]. Since
viscosity is a function of fluid temperature, the temperature
operating window (TOW) for a particular viscosity grade of
hydraulic fluid is a function of temperature. Figure 9 depicts

the TOW for straight grade mineral oil based hydraulic fluids. The viscosity grade indicated in the TOW corresponds to
ASTM D 2422, Classification of Industrial Fluid Lubricants
by Viscosity System. For example, ISO 32 hydraulic oil generally will provide satisfactory performance in a temperature
window of - 8 to 64C.
There are several methods for measuring the viscosity of
hydraulic fluid. The most widely utilized method is the ASTM
D 445, Standard Test Method for Kinematic Viscosity of
Transparent and Opaque Liquids. In this test, the time is
measured for a fixed volume of liquid to flow under gravity
through the capillary of a calibrated viscometer at a closely
controlled temperature. The kinematic viscosity is the product of the measured flow time and the calibration constant of
the viscometer. Based upon D 2442 and ISO 3448, the standard temperature for measuring hydraulic fluid viscosity is
40C [26]. Typically, the viscosity of a hydraulic fluid is 15-68
mm^/s (centistokes) at 40C. ASTM D 446, Standard Specifications and Operating Instructions for Glass Capillary Kinematic Viscometers, describes more than 15 types of viscometers that may be employed in performing a D 445 viscosity
test. With the exception of invert-emulsion type fluids, hydraulic fluids are generally transparent. Consequently, a tube
suitable for transparent liquids such as the popular CannonFenske viscometer may be used. For opaque liquids, a reverse-flow tube is required because it is difficult to see the
meniscus as the fluid flows by the timing marks on a standard viscometer. Cannon-Fenske tubes for viscosity measurement of transparent and opaque liquids are depicted in
Fig. 10.

1
MIXFn FN M

LUBRICA-|ON
B OUNDARY
LIJBRICATION
0.1

c
o
o
c
g>
o

0.01

^
O

o
0.001

0.001

0.01

359

0.1

hULL-hlLM
LUBRICATION
1
1

10

Sommerfeld number, (rjA// P) x 10"^


FIG. 8Stribeck Curve of coefficient of friction versus Sommerfeld Number (S), where
S = r}N/P. N shaft speed; P, average pressure between shaft and bearing due to applied
load; 7), lubricant viscosity.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

360 MANUAL 37: FUELS AND LUBRICANTS HANDBOOK


100

212

90

94 194

80

o
<D

176

84

70

158 LL.
140

73

60

64

5 50
-I

55

122

104 CO
CD
86 O.

44

I" 30

32

j5^20

CO

+10

10

+4
-2

-23

-4

-30 3 3

-22

-40
10

50

14

-15

-20

E
.<!>

32

-8

-10

CD
V

15

22

32

46

68

100

-40

ISO Viscosity Grade


FIG. 9^Temperature Operating Window (TOW) for 100 VI mineral oil based hydraulic fluids. Based upon survey of viscosity requirements for hydraulic pumps and motors, fluids
will generally provide satisfactory performance at the temperature range that corresponds
to 13 to 860 cSt.

FIG. 10Cannon-Fenske standard and reverse flow kinematic viscosity tubes, respectively.

Frequently in high-pressure hydraulic applications, the


loading conditions are sufficient to rupture hydrodynamic
lubricating films. Consequently, boundary and mixed-film
lubrication regimes play an important role in controlling
wear in hydraulic applications. In boundary lubrication, friction cind wear between two surfaces in relative motion are de-

termined by the properties of the surfaces and by properties


of the surfaces and lubricants other than viscosity [27 ]. In hydraulic equipment, these surfaces are typically composed of
ferrous or yellow metals. Under magnification, tribologicsJ
surfaces in hydraulic components reveal the presence of surface asperities. High load conditions cause these aperities to

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER
make contact, resulting in friction and weeir. In most cases,
mixed-film lubrication takes place and some hydrodynamic
lubrication occurs, even as "asperity contact" creates boundary conditions. Depending upon the extent of asperity contact, scuffing or adhesive wear may occur. A schematic description of the various wear processes specific to hydraulic
p u m p s is shown in Fig. 11. When cavitation, corrosion, or
scuffing wear processes generate particles that are the same
approximate size as p u m p clearances, synergistic wear may
take place. Sjmergistic wear ultimately leads to failure that
may appear to be abrasive in origin [28].
Wear protection under conditions of boundary lubrication
may be enhanced through the use of additives that interact
with surfaces to form protective chemical films. (See the Antiwear Performance Testing section for description of the
boundary lubrication additives utilized in hydraulic applications.) These chemical films reduce friction by decreasing the
shear strength of the surface relative to the underlying material. Thus, surface interaction under boundary lubrication
conditions is primarily between the low-shear strength chemical films rather than the metal substrate. Good wear protection and friction reduction result in enhanced equipment
durability, reduced heat generation, improved energy conservation, and many other operational advemtages.
Antiwear

Performance

13: HYDRAULIC

FLUIDS

pressures, and entry angles in a functioning hydraulic system


[29,30].
One of the more c o m m o n bench tests used for screening
the antiwear performance of a hydraulic fluid is the FourBall Method. There are two versions of the test for liquid lubricants: ASTM D 2783, Standard Test Method for Measurement of Extreme-Pressure Properties of Lubricating Fluids
(4-Ball Method) and ASTM D 4172, Standard Test Method
for Wear Preventive Characteristics of Lubricating Fluids (4Ball Method). The former method is generally used for evalu a t i n g extreme pressure gear lubricants while the latter
method is used for evEiluating antiwear hydraulic fluids. In
the 4-Ball Wear Test (D 4172), three half-inch diameter steel
balls are clamped together and covered with the lubricant to
be evaluated. A fourth ball of equal diameter is pressed with
a force of 1 5 ^ 0 kg into the cavity formed by the three stationary balls making a three-point contact (Fig. 12). Lubricants are evaluated by rotating the top ball under load at 1200
r p m for 60 min and measuring the average scar diameters
worn in the three lower balls.
In cooperative testing of fluids performed by members of
ASTM D02.L on Industrial Lubricants, the addition of zinc
dithiophosphate to 46 cSt mineral oil decreased the scar di-

Testing

The majority of hydraulic fluids are formulated with antiwear additives because surface loads associated with highpressure p u m p operation necessitate the use of fluids with
enhanced wear protection. There are a variety of test methods available for assessing the antiwear performance of hydraulic fluids. These tests may either be bench-top or fullscale tests employing high-pressure piston and vane pumps.
Bench tests are generally less expensive to perform t h a n
p u m p tests. However, translating bench test data into realworld performance can be problematic because of the complexity involved in simulating all of the materials, velocities.

(a)
FIG. 12The four-ball test:
(a) perspective view, (b) plan view.

EXTEERNAL
PAR"riCLE
INGREESSION
\Air:Ap

CAVITATION JElr

ASPERITY
CONTACT

FATIGUE
WEAR

ADHESIVE
WEAR

VVtArl

DEBRIS

WEAR
DEBRIS1

>

ABRASIVE
WEAR

TOTAL
WEAR

ELECTROLYTE
(WATER)

361

CORROSIVE
WEAR
WEAR

WEAR DEBRI S

DFRRI.q

FIG. 11Synergistic view of pump wear process. Fatigue, adhesive, and corrosive wear
can be triggered Independently. Resulting wear debris generation leads to abrasive wear.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

362

MANUAL

3 7: FUELS AND LUBRICANTS

HANDBOOK

ameter in 4-ball wear tests from 0.72 m m to 0.42 m m at 40kg


[31]. These results are tjpical of a mineral oil based antiwear
hydraulic fluid where average scar diameters of less than 0.50
m m are the norm (P. W. Michael, unpublished data).
While four-ball tests are effective in screening antiwear additive response, they do not directly correlate with p u m p
tests [32]. This is in part due to the fact that loads in the fourball tests are constant and do not pulsate in the same way
that a hydraulic p u m p does as sliding surfaces transition
from high pressure to low pressure regions of the pump. In
an effort to enhance the correlation between the four-ball test
and full-scale p u m p eveJuations Penn State University has
performed investigations involving sequential four-ball wear
tests. In the sequential four-ball test, wear scars are evaluated
at 10 and 40 kg and 600 r p m and the diameter of the scar is
measured after the fluid has been replaced by white oil in order to measure the durability of the antiwear film [33,34].
This method yields better correlation with vane p u m p tests.
The FZG Test is another bench test used for screening hydraulic fluids. FZG test equipment consists of two gear sets
arranged in a foursquare configuration (Fig. 13). The FZG
p r o c e d u r e is described in ASTM D 5182 S t a n d a r d Test
Method for Evaluating the Scuffing (Scoring) Load Capacity
of Oils. In this test, pre-examined gears are immersed in 1600

mL of oil that is heated to 90C (194F). The test gear set is r u n


in the test fluid for 15 min at successively increasing loads
until the failure criteria is reached. According to the ASTM
procedure, failure criteria are reached when the summed total width of scuffing wear damage from all 16 teeth is estimated to equal or exceed one gear tooth width. In DIN 51524,
Part 2, a m a x i m u m weight loss of 0.27 mg/kW h for antiwear
hydraulic oil is specified as well as a m i n i m u m damage stage
of 10. While Reichel reported a correlation between FZG Test
results and hydraulic fluid performance in vane pumps, correlation with piston p u m p performance has proven difficult
to establish [35].
The most widely referenced vane p u m p wear test for hydraulic fluids is ASTM D2882, Standard Test Method for Indicating the Wear Characteristics of Petroleum and NonPetroleum Hydraulic Fluids in a Constant Volume Vane
Pump (Vickers 104C). In this test, a hydraulic fluid is circulated through a rotary vane p u m p for 100 h at a p u m p speed
of 1200 r/min and a p u m p outlet pressure of 2000 psi. The
fluid temperature is controlled to 150F at the p u m p inlet for
most fluids. Petroleum based fluids with a viscosity greater
than 46 mm'^/s and some synthetic fluids must be evaluated
at 175F. At the end of the test, the total cam ring and vane
weight losses are measured and reported. Based upon ASTM

Drive gear case

Test gears with


long addenda

FIG. 13The Neimann (FZG) Four-Square Gear Test Rig.


Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER
D 6158, Standard Specification for Mineral Hydraulic Oils,
less than 50 mg of total wear is expected from properly formulated petroleum based antiwear hydraulic oil. For invertemulsion type fluids, higher wear rates in the 100-200 mg
range are c o m m o n while water glycol fluids routinely generate less than 50 mg wear in the D 2882 test.
While the D2882 test is a popular benchmark for evaluating hydraulic fluids, this method is not without its problems.
First of all, Vickers has discontinued production of the
V104C p u m p . This will ultimately necessitate the use of substitute hardware or abandonment of the test procedure. Second, rotor and bushing failures are common in the first few
hours of the test. This may be due to the fact that the p u m p
was originally designed for a m a x i m u m pressure of 1000
psig. Fluid performance in the V104C p u m p is evaluated at
1000 psi using the ASTM D 2271, Standard Test Method for
Preliminary Examination of Hydraulic Fluids (Wear Test). In
this procedure, the p u m p stand is operated for 1000 h, which
provides an extended evaluation of p u m p wear behavior under normal operating conditions. Xie et al. provide a detailed
discussion of the D 2882 Test Method in the Handbook of Hydraulic Fluid Technology [36].
For higher pressure a n d mobile applications Vickers
prefers their 35VQ25 vane p u m p for screening hydraulic
fluid wear performance (Table 5). In the 35VQ25 test, three
50-hour tests are conducted on the same charge of test oil.
For each 50-hour test a new p u m p cartridge is used. The test
rig is operated at 3000 psi and 200F with a p u m p speed of
2400 rpm. Vickers limits the amount of wear on each test kit
to 90 mg: 75 mg ring, 15 mg vanes. In addition there must be
no sign of scuffing on the cam ring.
The Denison T6C vane p u m p test is a variable pressure
vane p u m p test. In this test, a Denison T6CSH 020 p u m p cycles between 7 b a r (100 psi) and 250 bar (3600 psi) at onesecond intervals for 300 h [37]. The p u m p speed is nominally
1700 r/min a n d fluid t e m p e r a t u r e is m a i n t a i n e d at 80C
(176F) for mineral oil based fluids a n d 45C (113F) for
those based on water. The test is r u n in two 305-hour sequences. Each 305-hour test consists of a 5-hour break-in period followed by 300 h of high pressure cycling. After the first
305-hour test, the p u m p cartridge is removed for inspection
and a new cartridge is installed for the second sequence. The
second 305-hour sequence is r u n with 1% distilled water
added to the fluid. The first stage of the T6C test serves as an
aging mechanism and increases the susceptibility of the fluid
to the deleterious effects of water contamination. After the
second 305-hour sequence the p u m p cartridge is again removed for inspection. As with the 35VQ25 test, weight loss of
cam ring and vanes, vane tip profile, and visual appearance
of all components are all reported. In addition, a wet filterability test is performed on the fluid to determine if water
contamination will lead to filter blinding. (See the Filterability section for a discussion of filterability tests.)
Although the V104C and 35VQ25 vane p u m p tests have
served the industry well for many years, these tests are not
sufficient to screen hydraulic fluids that will be used in highpressure piston p u m p s applications [38]. Thus, piston p u m p
tests have been to qualify the antiwear capabilities of hydraulic fluids. Komatsu, Rexroth, and S u n d s t r a n d piston
p u m p tests are described below.
K o m a t s u developed a piston p u m p test to evaluate

13: HYDRAULIC

FLUIDS

363

biodegradable vegetable oil based hydraulic fluids [39]. This


test is based on a Komatsu HPV35+35 twin-piston p u m p using cycled pressure test conditions. In this test p u m p efficiency change, wear and surface roughness, formation of lacquer and varnish, a n d hydraulic oil deterioration are
evaluated.
Rexroth has proposed a three-stage piston p u m p test based
on the Brueninghaus A4VSO piston p u m p [40]. Stage one is
conducted at the m a x i m u m operating pressure and temperature and at the m i n i m u m viscosity specified for the fluid being tested. The test duration is 250 h at which time the p u m p
is dismantled and inspected. The second stage of the test is
pulsed pressure test at the m a x i m u m displacement of the
p u m p . This stage is operated for one million cycles. When
this stage is complete, the p u m p is dismantled and inspected.
The third stage is a variable displacement stage at maximum
pressure, maximum temperature, and m i n i m u m fluid viscosity. The test duration is 280 h at which time the p u m p is dismantled and inspected again. The final pass/fail assessment
is made with reference to a standard damage catalog.
The Sundstrand Water Stability Test Procedure test originally employed a Sundstrand Series 22 piston p u m p at a constant pressure [41]. Currently, this test procedure is conducted using a Sundstrand Series 90 piston p u m p with a 55-cc
displacement. The objective of the test is to determine the effect of water contamination on mineral oil hydraulic performance and yellow metal corrosion. However, other fluids, including water-containing fluids such as HFB and HFC fluids,
may also be evaluated using this test. The test duration is 225
h, at which time it is disassembled and inspected for wear,
corrosion, and cavitation. If the flow degradation is equal to
or greater than 10%, the test is considered to be a "fail."
Antiwear

and Extreme

Pressure

(EP)

Additives

Antiwear and EP additives prevent wear of metal surfaces by


forming a protective chemical film between moving parts.
These additives have traditionally been labeled as antiwear or
extreme pressure (EP), depending on the mechanism of protection. Antiwear additives are generally considered to form
protective films that adsorb on the metal surface and function effectively under relatively mild conditions of load and
t e m p e r a t u r e . Extreme pressure additives form protective
films by reacting with the metal surfaces at localized high
temperatures to form low shear strength films that are relatively insoluble in the bulk oil. In either case, tribological
contact is between the surface films rather than the metals.
Various types of chemistry are employed in the prevention
of wear in hydraulic applications. Typical compounds include zinc dialkyldithiophosphates (ZDTP), tricresylphosphates (TCP), sulfur compounds, amine phosphates, dithiocarbamates, and other chlorinated, phosphorus/sulfur, and
molybdenum compounds.
Water Content a n d Hydrolytic Stability
In many hydraulic systems, the lubricant is susceptible to
contamination with water. Contamination with water can
lead to a host of problems including loss of lubricity, corrosion, additive degradation, and filter plugging. Consequently,
machine builders and equipment users often attempt to limit
the amount of water that enters their hydraulic systems. At
the same time, fluid formulators endeavor to manufacture

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

364 MANUAL 3 7: FUELS AND LUBRICANTS HANDBOOK


TABLE 5Machine builder specifications for antiwear hydraulic oil.
Properties
Method(s)
ISOVG
Kinematic Viscosity, cSt
D445
0C max., calc. D 5133
40Cmax.
40C min.
100Cmin.
Flash Point C min.
D92
Fire Point, C min.
D 92
Pour Point, C, max
D97
Color, max
D 1500
ISO Contam. Code, max
ISO 4406
Density @15C
D1298
TAN, mg KOH/g, max
D664/ D974
Rust Test A
D665 A
Rust Test B (Salt Water)
0665 B
Cu Rating (3 hr, 100C), max. D130
TOST Oxidation, Hours to 2.0 ,^^,
TAN
"*^
Air Release @ SOX, minutes Q J ~ ,
(max)
Foam tendency/statxiity
D892
Seq 1
max
Seq II
max
Seq III
max
DemulsilJility @ 54C
D1401
FZG Fail Stage
D 5182
Change in Hardness
NBR1,168hrs@100C
Change in Volume (%)
NBR1,168hrs@100C
Viscosity Index, min
D 2270
Aniline Point C min.
D 611
CM Thermal Stabtity
D 2070 A
Viscosity Change, % max
TAN Variation, % max *
Comparative IR Scan
Sludge, mg/100 ml max
Cu metal removed, mg/200 ml, max.
Copper rod appearance, rating (max.)
Steel deposits, mg/200 ml, max
Steel metal removed, mg/200 ml, max
Steel rod appearance, rating (max)
Oxidation (1000 h)
D4310
AN, mgKOH/g max
Total sludge, mg max.
Copper, mg max
Iron, mg max
Hydrolytic Stability
D 2619
Copper wt loss, mg/cm^ max
Water layer TAN. mgKOH max
V104 C Pump
mg wear, max
D 2882
Vickers 35 VQ 25 Pump Test
Vane Wear, mg max
Rir^ Wear, mg max
Denison P-46 (100 h)
DenisonTBC, vane wear
TP-30283
Cam ling wear
Denison Fnterability Test, sec TP 02100
Dry, max
Wet, max

Denison HF-0

Vickers

Requirements

Requirements

6 M LS-2

Cincinnati iWachine
.

P68
32

P70
46

P69
68

35.2
28.8

50.6
41.4

74.8
61.2

188
215

196
218

196
218

LH-02
22

LH-03
32

LH-04
46

LH-06
68

300
24.2
19.8
4.1
175

420
35.2
28.8
5
190

780
50.6
41.4
6.1
190

1400
74.8
61.2
7.8
195

-21

-18

-15

-12

10

10

19/16/13
0.84 - 0.90
<1.S>
Pses

Pass
Pass

Pass
<1b>
<1500>
5

<:50/0>
<SO/0>
<50/0>
Timeto40/40)(O/W/E)
<30>
<10>

-10
-/O
-10

90
100

<90>

Oto-7

Oto-7

Oto-6

OtolS

0to12
<95>

0to12

OtolO

<5>
<50>
Record
< 25 mg. /100ml >
<10>
<5>

<5>
<50>
100
10
Report

Oto-8

< 25 mg. /100ml >


<10>
<5>
<3.5>
<1>
<1.5>

<1.5>

2
200
50
SO
(1)
0.2
4.0

<0.2>
<4>
I-2S6-S
SO
M-2950^
15
75

< 50 >

(2)

Satisfactory
Satisfactoiy

600
2xdry

(1) Rqmnt. Sut>ject to Denison discretion (based on other pump/fietd history)


(2) D 2882 mn at 79,4C (higher temp.) for ISO 68 and higher grade.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

<10>
<50>
no smear, scratch, etc
< 0.01 >
No distress
<600>
< 2 X dry >

CHAPTER 13: HYDRAULIC FLUIDS


hydraulic fluids that resist chemical degradation or hydrolysis in the presence of water and heat. Several ASTM methods
are used to monitor water content of hydraulic fluids as well
as their ability to resist hydrolytic degradation.
Distillation, Centrifuge and Karl Fisher Titration Tests
In ASTM D 95, Standard Test Method for Water in Petroleum
Products and Bituminous Materials by Distillation, the material to be tested is diluted with a water-immiscible solvent
such as toluene and heated under reflux conditions. The resulting distillate is condensed and separated in a trap. The
amount of water present in the sample is determined by observing the volume of water settled in the graduated section
of the trap.
Centrifuge tests such as ASTM D 96, Standard Test Method
for Water and Sediment in Crude Oil by Centrifuge Method,
can also be used for un-emulsified or insoluble water contamination in fluids. While distillation and centrifuge methods provide reasonably accurate results for samples that contain free water contamination, these methods are generally
not sensitive enough for hydraulic applications. A more accurate method for quantifying water in hydraulic fluid is the
Karl Fischer test (ASTM D 1744, Standard Test Method for
Determination of Water in Liquid Petroleum Products by
Karl Fischer Reagent) [42]. In this test, the fluid is dispersed
in a solvent such as methanol and titrated with standard Karl
Fisher reagent to an electrometric endpoint (Fig. 14). The
endpoint of the titration, at which free iodine is liberated.

365

may be registered either potentiometricly or by color indication. Although this method has the capability to be more accurate than distillation or centrifuge techniques, the Karl
Fisher Test is susceptible to chemical interference. Calcium
sulfonate, magnesium sulfonate, ZDTP and other oil additives react with iodine and have been known to interfere with
the titration [43].
Hydrolytic Stability Testing
Hydrolytic stability refers to the lubricant's resistance to
chemical interactions with water that result in undesirable
changes to fluid properties. Certain chemical components
may react with water to decompose or form undesirable
byproducts of hydrolysis. Heat and catalysts such as copper
can accelerate the process of hydrolysis. Hydrolytically unstable oils form insoluble contaminants and acidic compounds that create hydraulic system malfunctions similar to
those produced by oxidation and thermal degradation of fluids. Furthermore, antiwear additives and corrosion preventatives that are susceptible to hydrolysis are likely to lose their
ability to perform their critical functions in the presence of
heat and water.
ASTM D 2619, Standard Test Method for Hydrolytic Stability of Hydraulic Fluids (Beverage Bottle Method) is used to
measure this fluid property. In this test, 75 g of fluid and 25
g of water are sealed in a beverage bottle with a copper strip.
The test bottie is rotated in an oven for 48 h at 93C (200F).
At the end of the test, the oil and water layers are separated

FIG. 14The Karl Fisher apparatus (a) titrant solution, (b) burette, (c) titration cell with
electrode, (d) solvent, (e) waste.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

366

MANUAL

3 7 ; FUELS AND LUBRICANTS

HANDBOOK

and insolubles are weighed. Viscosity a n d acid numbers are


also determined. Based upon the Denison HF-0 specification
(see Table 5 for details of this specification), t h e weight
change of t h e copper specimen should b e less t h a n 0.20
mg/cm^ a n d the water layer acidity should be less than 4.0 mg
KOH. Since exposure to water can be expected throughout
the life of a fluid, hydrolytic stability is a n important design
characteristic of hydraulic fluids.
In genereJ, there are no additives specifically used to improve hydrolytic stability. Instead, hydrolytic stability is
achieved by appropriate selection of stable components that
maintain effectiveness even in t h e presence of water. Hydrolytic stability is also a key factor in the wet filterability behavior of hydraulic oils (see the Filterability section) [44].
Demulsibility
Demulsibilty is the term used to describe a fluid's ability to
separate from water. As discussed above in the Water Content
and Hydrolytic Stability section, water contamination of the
hydraulic oil may lead to various problems that adversely affect both fluid a n d equipment durability. Thus, it is desirable
for hydraulic oil and water to separate as quickly as possible.
In many industricJ applications, water is drained from the
hydraulic oil reservoirs as it separates and settles on the bottom. For fluids with poor demulsibility, the separation is either very slow or unlikely to occur to any significant degree.
Demulsibility

Testing

levels, entrained air is visible to the h u m a n eye as larger bubbles and can cause the oil to become cloudy. Uncontrolled air
contamination results in a n u m b e r of undesirable consequences. Entrained air increases the compressibility of the
fluid and can adversely affect its response to hydraulic control mechanisms or devices, especially in high-pressure systems. Dissolved or entrained air expands into larger bubbles
as its solubility in the fluid decreases as a result of exposure
to vacuum conditions at the p u m p inlet. This leads to noise
and cavitation, which is the dynamic process of gas cavity
growth a n d collapse in a liquid [47]. Several studies of this
p h e n o m e n o n have suggested theoretical m e c h a n i s m s a n d
documented experimental evidence of wear a n d increased
oxidation due to cavitation [48].
Foaming is very much rooted in the fundamentcJ problem
of air contamination and consequently, results in many of the
same negative effects of air entrainment. It is characterized
by the formation of a mass of relatively large bubbles on the
surface of the fluid and is usually brought about by turbulent
return of oil to the reservoir or migration of entrained air to
the surface. It is desirable to have fluids with a low tendency
to form foam in the first place a n d have the foam collapse
quickly once formed. For effective foam control, the rate of
foam collapse must be faster t h a n the rate at which entrained
air migrates to the surface to form the foam. Otherwise, the
foam layer will continue to increase and air may eventually
be re-dispersed in the bulk fluid [49]. In severe cases, oil that
produces a significant amount of foam may bubble out of hydraulic reservoir breathers, creating a fluid spill.

The speed at which water is separated from oil and the tendency of an oil to form a cuff of emulsified oil at the interface
between the oil and water phases may be measured by ASTM
D 1401, S t a n d a r d Test Method for W a t e r Separability of
Petroleum Oils and Synthetic Fluids. In this test, a 40 ml sample of oil a n d 40 ml of distilled water are stirred for 5 min at
54C (130F) in a graduated cylinder. The time required for
the emulsion to separate into water a n d oil phases is
recorded. An oil with good demulsibility will completely separate in 30 m i n or less without a "cuff' of emulsified oil between the phases [45].

Air entrainment has increasingly become a concern due to


a trend toward smaller reservoir sizes. Shorter fluid residence times therefore dictate use of hydraulic fluids with improved air release characteristics for the reasons discussed
above. Several studies have shown that fluid viscosity is a
critical factor influencing air release properties. Within a
given class of fluids, higher viscosity and lower oil temperatures translate into slower air release characteristics. While
different classes of base fluids have demonstrated unique air
release advantages, there has been little success in identifying
additives that improve air release properties of a base fluid.

Demulsifiers

Foam and Aeration

Demulsifiers are chemicals used to alter the surface tension


at the oil/water interface to accelerate separation. T3rpical
demulsifiers include alkylphenol ethers, low molecular
weight synthetic sulfonates, and polyoxyalkylate resins.

Because of the importance of properly managing air contamination in hydraulic fluids, there are a n u m b e r of standardized test methods for evaluating this feature of fluid performance. The foaming tendency a n d stability of oil may be
measured by ASTM D 892, Standard Test Method for Foaming Characteristics of Lubricating Oils. In this test, an oil
sample is equilibrated at 24C (75F). Air is bubbled through
oil for 5 min, and then the oil is allowed to settle for 10 minutes. The volume of foam is measured at the end of both periods. The test is repeated at 93.5C (200F) and again at 24C
(75F) after the foam breaks. Various levels of foaming tendency are permitted by industry standards, but stable foam is
generally not tolerated [50,51].
Not only must a hydraulic fluid resist the tendency to form
stable foam, it also must allow air to rapidly rise and separate
from the fluid. The Waring blender test is one test method
that may be used to measure the air release properties of fluids [52]. In ASTM D 3519, Standard Test Method for Foam in
Aqueous Media (Blender Test), 200 ml of the fluid is stirred

Aeration a n d Foam
Under normal conditions there is always air present in a hydraulic fluid. By volume, it is present at about 7-9% at room
temperature a n d atmospheric pressure [46]. In this state, it is
not visible to the h u m a n eye and thus referred to as dissolved
air. Higher temperatures and/or lower pressures (such as vacu u m conditions) lead to lower dissolved air levels. (See chapter on compressor lubricants for detailed discussion on gas
solubility and methods of measuring gas solubility.)
Fluid circulation through hydraulic systems and reservoirs
may cause mecheinical introduction of air into hydraulic fluids, particularly if reservoir size or design does not allow sufficient residence time for air separation to occur. At elevated

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

Tests

CHAPTER
at an agitation rate of 4000 to 13000 r p m for 30 s. The meixim u m total height at zero time, at 5 m i n a n d 10 m i n is
recorded in order to assess the foaming and aeration tendency of a fluid under high shear conditions.
Air release properties of a hydraulic fluid may also be
quantified by IP 313, DIN 51381 or ASTM D 3427, Standard
Test Method for Air Release Properties of Petroleum Oils. In
these tests, the time in minutes for finely dispersed air in oil
to decrease to 0.2% under standard test conditions is measured using a density balance. Air release times and specifications typically vary with oil viscosity.
Defoamants
Antifoam additives, generally referred to as defoamers or defoamants, are materials that destabilize the liquid film that
surrounds air bubbles. The most commonly used defoamants
are silicone polymers (particularly polydimethylsiloxanes),
which function as finely dispersed marginally soluble liquid
particles. Since silicon defoamants have very low surface tensions, they tend to accumulate at air/oil interfaces. When the
larger bubbles rise to the surface and join other bubbles to
form foam with only very thin films separating them, silicone
defoamants cause these films to rupture, thus accelerating
collapse of the foam. While silicone defoamants reduce the
foaming tendency of a fluid, they may also tend to increase
air entrainment (Fig. 15) [53].
Besides affecting air entrainment in hydraulic fluids, silicone defoamants tend to have poor filterability and storage
stability due to their marginal solubility in oil. Non-silicone
defoamants are increasingly used to address these disadvantages. Polyalkylacrylate additives are the most common class
of non-silicone defoamants recognized in the industry. Although they do not possess the disadvantages of the silicone
types, these polyalkylacrylates must be used at higher concentrations to deliver equivalent performance.

13: HYDRAULIC

It is widely recognized that beyond proper fluid selection,


good fluid maintenance is the key to reliability and durability
of hydraulic equipment. Fluid maintenance is closely linked
to fluid cleanliness and filtration. Filtration devices, therefore, are critical tools for maintaining hydraulic fluids and
system components. Hydraulic fluid "filterability" is concerned mainly with the appropriate flow characteristics of
the fluid through filter media. For proper operation, the fluid
should readily flow with m i n i m u m pressure drop across the
filter and with negligable depletion of additives. The viscosity
and chemistry of the lubricant will affect filterability. Therefore, filter size and materials should be compatible with the
circulating fluid. The drive to increase hydraulic system reliability through the use of fine filtration magnifies the importance of this performance parameter.
Filterability

Tests

Due to the likelihood of water contamination in many hydraulic systems and its potential impact on fluids, most of the
filterability tests are designed to r u n dry and wet (with water
added). Hydraulic fluid filterability tests generally consist of
filtering a specified quantity of fluid t h r o u g h a standard
medium while monitoring changes in flow rate (Table 6). The
results are tj^pically reported in terms of a ratio between flow
rates with and without water added to the fluid. This approach attempts to account for changes in filterability behavior independent of viscosity. In Denison TP 02100 the
time required for complete flow of a standard volume of fluid
through a specified filter is evaluated. In the Pall Filterability
Test the differential pressure across a specified filter assembly is monitored over the duration of the test and cin appropriate limit is established to discriminate between fluids with
good and poor filterability behavior. While key equipment

OIL WITH
SILICONES

O
O
eg
<

h-

SETTLING OR
"TRANQUIL PHASE"

BLOWING OR
TURBULENT PHASE"

367

Filterability

AIR RELEASED DURING


BLOWING PHASE

VOLUME OF AIR
BLOW IN

FLUIDS

>

TIME
FIG. 15Impact of silicone defoamer on foaming tendency and air release. Silicone defoamer decreases the tendency of the oil to generate foam while increasing the tendency
of the fluid to retain air below its surface.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

368

MANUAL

37: FUELS AND LUBRICANT

Method

TABLE 61Filterability tests.


AFNOR
Pall"
0.8jU.M
0.2
70 h
70C

Medium pore size


Percent water added
Aging time
Temperature

HANDBOOK

Denison

3/LiM
1.0
24 h
70C

1.2 ^M
2.0
None
25C
"Parkhurst, H., Pall Filterability Index Test for Paper Machine Oils, SLS Report No. 5669, April 1995.

builders and industrial manufacturers may require fluids to


meet certain filterability criteria as measured by these tests,
global hydraulic oil specifications (i.e. ASTM D 6158, ISO
11158, DIN 51524) have not yet incorporated these procedures.
Filterablility

Additives

From a formulation standpoint, identifying and replacing additives with potential filterability problems (i.e., filter material incompatibility, gel-forming tendency, hydrolytic instability, etc.) has been the primary method of improving fluid
filterability. Recently, dispersants have been identified that
enhance filterability by preventing agglomeration of insoluble species present in the fluid. These dispersants are typically alkyphenol-based or alkyl succinimide polymers of
varying molecular weights.

Corrosion Protection
Chemical contaminants and corrosive by-products of fluid
degradation can cause surface attack of metallic hydraulic
system components. Ferrous metal corrosion in a hydraulic
system is most often caused by water contamination, while
copper and its alloys are susceptible to attack by the products
of high temperature fluid degradation. Rusting of ferrous
metal is an electrochemical reaction that occurs between the
parent metal and the thin oxide layer on the metal surface
formed as a result of exposure to the atmosphere [20]. Rust,
which is hydrated iron oxide, compromises the integrity of
the metaJ surface and adversely affects other important fluid
properties w h e n it contaminates the bulk fluid. Ferrous
metal corrosion protection in hydraulic systems is usually accomplished by incorporating surface-active additives such as
rust inhibitors. There are several ASTM methods for evaluating the corrosion inhibition properties of hydraulic fluids.
Corrosion

and Rust

Testing

The ability of fluids to prevent rusting of ferrous parts due to


water c o n t a m i n a t i o n m a y be m e a s u r e d by ASTM D 665,
Rust-Preventing Characteristics of Inhibited Mineral Oil in
the Presence of Water. In Part A of this test, 10% distilled water is added to oil that has been heated to 60C (140F).
Round steel rods are polished to remove their oxide coating
and immersed in the oil. The oil-water mixture is continuously stirred to avoid separation while the temperature is
maintained at 60C. At the end of 24 h the specimens are inspected for rust (Fig. 16).
In Part B of the method, the same procedure is used, except
synthetic seawater is substituted for distilled water. As described in Part B, synthetic seawater is made by the addition

of sodium chloride, magnesium chloride, calcium chloride,


and several other ionic compounds to distilled water. Part B
is particularly pertinent to maritime hydraulic fluid applications where seawater, rather t h a n fresh water or condensation, is a likely source of contamination.
The standard test method for measuring vapor phase corrosion inhibition of hydraulic fluids is ASTM D 5534, Test
Method for Vapor-Phase Rust-Preventing Characteristics of
Hydraulic Fluids. In this test, a steel specimen is attached to
the cover of an ASTM D 3603 test apparatus that contains hydraulic fluid maintained at a temperature of 60C (140F).
ASTM D 3603 is the Horizontal Disk Method for Rust-Preventing Characteristics of Steam Turbine Oils in the Presence
of Water. The specimen is then exposed to water and hydraulic fluid vapors for a period of 6 h. At the end of this time,
the specimen is inspected for evidence of corrosion and results are reported on a pass-fail basis. The ASTM D 5534 test
is particularly relevant for water-glycol and invert-emulsion
hydraulic fluids because corrosion of the underside of reservoir covers has been observed in systems that use these fluids.
Accelerated corrosion can also occur when dissimilar metals are in electrical contact in the presence of an electrolyte
(i.e., conductive solution). This corrosion mechanism,
known as galvanic corrosion, has been found to be particularly relevant for certain biodegradable oils [54], The ability
of a fluid to prevent galvanic corrosion may be measured by
ASTM D 6547, Test Method for Corrosiveness of a Lubricating Fluid to a Bi-Metallic Couple. In this test, a brass clip is
fitted to the oil coated surface of a steel disk. The bi-metallic
(brass/steel) couple is then stored in 50% relative humidity
for ten days. At the end of the ten-day period, the surfaces are
inspected for evidence of staining like that depicted in
Fig. 17. The steel disks are rated on a pass-fail basis.
Sulfur containing additives such as zinc dithiophosphate,
sulfurized olefins, organic polysulfides, and carbamates may
be used as antiwear and extreme pressure additives in hydraulic fluids [55]. Depending u p o n the chemical activity of
these sulfur compounds, hydraulic fluids exhibit varying degrees of corrosiveness to copper when activated by high temperatures. ASTM D 2070, Standard Test Method for Thermal
Stability of Hydraulic Oils is one of the most effective methods for predicting the corrosiveness of a hydraulic fluid to
copper and its alloys. The ASTM D-2070 test measures the
aggressiveness of chemical constituents in the fluid toward
yellow metals when aged under high temperature conditions.
(See the section on High Temperature Oxidation Tests)
In some cases, such as when a hydraulic fluid is contaminated with sulfur containing metalworking fluid, the fluid
may exhibit corrosivity to copper without requiring thermal
degradation. The standard test method for measuring the copper corrosion properties of oil is ASTM D 130, Standard Test
Method for Detection of Copper Corrosion from Petroleum
Products by the Copper Strip Tarnish Test. In this test, a polished copper strip is immersed in oil and heated for a predefined period of time. At the end of the test, the copper strip's
appearance is compared to a standard. The rating system used
for the D 130 test appears in Table 7. The rating system is on
a scale of one to four. The higher the copper strip rating, the
greater the degree of copper corrosion. Color standards are
also available from ASTM for rating copper strips [56].

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER 13: HYDRAULIC FLUIDS

369

FIG. 16ASTM D 665 passing vs. failing rod.

FIG. 17Galvanic corrosion: staining on test specimen by vegetable oil.

Corrosion Inhibitors, Rust Inhibitors, and Metal


Passivators
Corrosion Inhibitors, Rust Inhibitors, and Metal Passivators
are designed to prevent deterioration of metal surfaces that
are in contact with the lubricant. Corrosion inhibitors are polar molecules that are surface active. They adsorb on the
metal surface and inhibit the electrochemical reaction that
produces rust. Some hydraulic fluids, particularly those used
in applications that require enhanced fire resistance, are for-

mulated with water. Such fluids have entirely different corrosion inhibition requirements. For instance, water glycol
hydraulic fluids must prevent corrosion in the vapor phase
above the liquid due to evaporation. Thus they are formulated with vapor phase corrosion inhibitors such as morpholine. Typical classes of rust inhibitors include metallic sulfonates, amine phosphates, simple fatty acids, and succinic
acid esters. Triazoles, or derivatives thereof, are commonly
used metal passivators.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

370

MANUAL

3 7: FUELS AND LUBRICANTS

HANDBOOK

TABLE 7Copper strip classifications.


Rating

Designation

la

Slight tarnish

lb
2a
2b
2c

Slight tarnish
Moderate tarnish
Moderate tarnish
Moderate tarnish

2d
2e
3a
3b

Moderate tarnish
Moderate tarnish
Dark tarnish
Dark tarnish

4a

Corrosion

4b
4c

Corrosion
Corrosion

Light orange, almost the same as


freshly polished strip
Dark orange
Claret red
Lavender
Multicolored with lavender blue or
silver overlaid on claret red
Silvery
Brassy or gold
Magenta overcast on brassy strip
Multicolored with red and green
showing (peacock), but no gray
Transparent black, dark gray or
brown with a trace of peacock
Graphite or lusterless black
Glossy or jet black

Seal Compatibility
Very critical to the successful operation of a hydraulic system
is the ability to prevent leakage and accidents that are a result
of failed seals. Leaks can lead to contamination, loss of pressure, loss of lubricating fluid, and environmental damage depending on the severity of the spill. In extreme temperature
and pressure operations, sudden failure of seeds may have life
threatening consequences, considering the potential for explosions, fires, etc. [57]. Hydraulic fluids and elastomeric
seals are composed of complex chemical components that
can interact as they come into contact. Depending on the
chemistries involved, time, t e m p e r a t u r e , a n d mechanical
stresses cause fluid interactions with the seal material, resulting in swelling or shrinkage of the elastomer compound.
It is desirable to select seal materials that exhibit minimal
change in hardness, volume, tensile strength etc. in service.
Slight swelling of seals is preferable to shrinkage as indicated
in Table 8. This is because a reduction in seal volume may result in leakage of fluid due to failure of the seal to fill the
gland that retains it in place.
Seal Compatibility

TABLE 8Recommended property change limits for


determining compatibility of elastomer seals for
industrial hydraulic fluid applications.

Description

Testing

In general, industry recognized seed compatibility tests entail


exposure of the elastomer material to the test fluid for a specified duration and at a standard temperature u n d e r static
conditions. Familiar industry seal compatibility tests include
ISO 7619, ISO 6072, DIN 53 538, and ASTM D 6546-00, Standard Test Methods for and Suggested Limits for Determining
Compatibility of Elastomer Seals for Industrial Hydraulic
Fluid Applications. Other major organizations such as ASTM
and SAE also have related specifications for sealing devices.
Due to variations in elastomer chemistry, it is necessary to
perform compatibility tests on the specific materials being
used. While most standard tests measure changes in hardness, stress/strain properties, and volume changes after exposure to the test fluid, translation of these results to a practical application m a y be difficult, since geometry and
mechanical conditions of the targeted application profoundly impact the elastomer. It is therefore recommended
that seal materials be tested u n d e r conditions that closely
simulate the actual application [58].

Maximum
Volume
Swell,

Time in
Hours
24
70
100
250
500
1000

Seal Swell

Maximum
Vol.
Shrinkage,

Hardness
Change,
Shore A
Points

15
15
15
15
20
20

-3
-3
-3
-4
-4
-5

7
7
8
8
10
10

Maximum
Tensile
Strength
Change, %
-20
-20
-20
-20
-25
-30

Agents

These chemicals react with the elastomeric materials to


cause slight swelling or softening to counteract the typical effects of temperature and mechanical stress. Seal swell agents
are typically used with base fluids having very low aromatic
content. Aromatic derivatives or phosphate esters are typically used to enhance the seal swell characteristics of a fluid.
Coolant Separability
Hydraulic systems used in machine tool operations are susceptible to contamination by aqueous cutting fluids, which
contain components with poor oxidation resistance, high deposit forming tendency, and/or high corrosivity. In metcdworking applications, the hydraulic fluid may be considered
a contaminant of the cutting fluid that alters its effectiveness
in metal removal operations. Regardless of the perspective, a
mix of these two categories of fluids is undesirable, especially
if they have not been designed to be compatible. In this case,
compatibility is defined as the ability of either fluid to complement, enhsince, or at least have no impact on the performance of the other w h e n mixed. The lubricant's ability to
readily separate from coolants is highly desirable in most
cases. However, the variety and complexity of coolant
chemistries makes it difficult to ensure good separability of
the hydraulic oil from all metalworking fluids [59].
There are generally n o additives specifically designed to
improve coolant separability, since coolant chemistries vary
so widely. The t3?pical approach is to formulate a lubricant to
have good demulsibility (water separability) and then test its
compatibility with specific coolants with which it is expected
to come into contact.
Coolant

Separability

Testing

A standard industry test method for assessing lubricant compatibility with coolants has not yet been established. However, some Icirge industrial manufacturers and lubricant suppliers do have in-house test procedures designed to simulate
oil contamination by a low percentage of coolant, as well as
coolant contamination by a low percentage of oil (typically
referred to as tramp oil). In general, these procedures consist
of mixing the lubricant with the coolant at a specified ratio
a n d t e m p e r a t u r e for a s t a n d a r d duration. The fluid container, t5^ically a graduated cylinder, is then allowed to sit
while the degree of separation between the coolant and the
lubricant is observed at specific time intervals. Properties
such as additive leaching and foam stability may also be ob-

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER
served. Rapid separation, implying absence of a stable emulsion or cuff (the layer between way oil and coolant) at the interface, is very desirable (Fig. 18).
Shear Stability
Mobile hydraulic equipment such as excavators, farm tractors, cranes, and timber harvesters frequently are required to
operate under extreme high and low temperature conditions.
To accommodate wide-ranging environmental conditions,
hydraulic fluids with enhanced viscosity - temperature properties are often employed. These fluids t3^ically contain viscosity index improving polymers that thicken oil at high temperatures, while having little impact u p o n their low
temperature fluidity. Viscosity index (VI) is a common means
for expressing the variation of viscosity with temperature.
The viscosity index of an oil is calculated from the measured
viscosity of the fluid at 40 and lOOX using ASTM Method D
2270, Standard Practice for Calculating Viscosity Index from
Kinematic Viscosity at 40 and 100C. A high VI indicates less
relative change in viscosity for a given change in temperature. Vl-improved oils are commonly referred to as multigrade oils, because they meet both the low temperature requirements of low viscosity oils and the high temperature
requirements of higher viscosity oils. Conceptually, an SAE

Good

13: HYDRAULIC

FLUIDS

371

lOW-30 multigrade oil consists of a lOW base oil and sufficient polymer to thicken the oil at 100C to a viscosity equal
to that of an SAE 30 weight oil (Fig. 19).
Viscosity Index Improvers are typically subjected to mechanical degradation due to shearing of the molecules in high
stress areas such as between gear teeth in gear pumps and
vane-ring interface in vane p u m p s . High pressures generated
in hydraulic systems subject fluids to shear rates up to 10^ s~'
[60]. Not only does hydraulic shear cause fluid temperature
to rise in a hydraulic system, but shear may bring about permanent viscosity loss in hydraulic fluids [61]. Permanent viscosity loss results from mechanical scission of polymer
molecules in multigrade hydraulic fluids and often occurs after a relatively short period of time (<24 hours of operation).
The polymer (as opposed to the base oil) is susceptible to mechanical shear because it has a higher molecular weight and
therefore a larger molecular volume. As a result, with polymer-containing multigrade hydraulic fluids, the functional
viscosity of an oil may differ from that predicted from kinematic viscosity measurements of new oil [62].
Shear Stability

Tests

It is desirable to formulate hydraulic fluids with shear-stable


VI improvers so that the fluid retains its viscosity properties
throughout its service life. Several laboratory test methods are

Fair

FIG. 18Good vs. Bad coolant separability.


Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

Poor

372 MANUAL 37: FUELS AND LUBRICANTS

HANDBOOK

Multigrade Oils
100,000
10,000
1,000

of resistance to mechanical shear, as well as their chemical


and thermal stability. For a given polymer type, shear stability decreases with an increase in molecular weight. Shear is
indicated by a loss in fluid viscosity (Fig. 20). The "thickening
efficiency" of the viscosity modifiers generally increases with
an increase in molecular weight for a given polymer type. Selection of the best VI Improver must entail consideration of
viscosity requirements, shear stability, and thermal and oxidative stability in actual equipment operation.
Low Temperature Pumpability

2-i
-40

.
^40
100
150
Temperature, C
FIG. 19Impact of VI improver on lubricant viscosity.
-20

designed to stress multigrade oils so that they produce a permanent viscosity loss such as would take place in service. The
two methods generally used are mechanical shearing with a
Bosch diesel fiiel injection pump and sonic shearing with a
high frequency sonic oscillator. In ASTM D 6278, Test Method
for Shear Stability of Polymer Containing Fluids Using a European Diesel Injector Apparatus, the polymer-containing
fluid is passed through a diesel injector nozzle at a shear rate
that causes polymer molecules to degrade. Under standard
test conditions, the kinematic viscosity of the fluid is measured after 30 to 250 cycles through the injector pump to determine the extent of permanent viscosity loss that has taken
place. In ASTM D 5621, Standard Test Method for Sonic
Shear Stability of Hydraulic Fluid, the polymer-containing oil
is irradiated with a sonic oscillator for 40 min and changes in
kinematic viscosity are measured. Based upon data from
Kopko and Stambaugh, the Fuel Injector Shear Stability Test
lacks the necessary severity to predict permanent viscosity
loss produced by hydraulic equipment [62]. However, 40 min
of irradiation with a high frequency sonic oscillator produced
viscosity changes that closely correlate to that experienced in
the ASTM D 2882 Vane Pump Test. Consequently, this test
method has become the basis for ASTM D 6080, Practice for
Defining the Viscosity Characteristics of Hydraulic Fluids.
Viscosity Index Improvers
Viscosity Index Improvers (also referred to as viscosity modifiers) are high molecular weight poljTuers that reduce the
magnitude of viscosity change as a function of temperature.
They function by enabling the oil to retain thickness at higher
temperatures while having minimal impact on viscosity at
lower temperatures. In general. Viscosity Index Improvers
are oil-soluble organic polymers with molecular weights
ranging from about 10000 to 1 million. The oil temperature
controls coiling of the polymer molecules, which in turn controls the degree to which the polymer increases viscosity. The
higher the temperature, the less the coiling and the greater
the "thickening" effect of the polymer. Therefore, as temperature increases, there is less thinning of the lubricant compared to non-polymer-containing oils.
The performance of VI Improvers is also described in terms

Paraffinic mineral oils, which comprise the bulk of hydraulic


fluids, contain some amount of wax that forms crystalline
structures at low temperatures. As these structures form, the
oil becomes more viscous. At very low temperatures the fluid
may become gel-like or even solid. For hydraulic systems,
poor low temperature flow characteristics can result in catastrophic failures. During start-up at very low temperatures, significant pump cavitation can occur due to inadequate oil flow.
Low Temperature Pumpability Tests
A number of bench tests are commonly used to evaluate low
temperature flow characteristics of lubricants. One of the
most common tests specified for this purpose is ASTM D 97,
Standard Test Method for Pour Point of Petroleum Products,
which measures the lowest temperature at which a lubricant
will flow. ASTM D 6351, Test Method for Determination of
Low Temperature Fluidity and Appearance of Hydraulic Fluids is used for evaluating the pour characteristics of
biodegradable oils. While this test gives an indication of low
temperature flow characteristics of the fluid, it does not necessarily address fluid performance in many applications subjected to very low temperatures. Pumps, motors, engines, and
many types of lubricated machinery require that the lubricant be pumped or circulated effectively at start-up. As a result, several tests have been developed to determine fluid viscosity at low temperature. For hydraulic systems, tests such
as Brookfield (ASTM D2 983), Scanning Brookfield SBV
(ASTM D 5133), and Mini Rotor Viscometer MRV (ASTM D

Multigrade Oils and Shear Stability


Quiescent Polymer Coil in
Oil Solution

Reversible/
Orientation of Coil
Under Shear Forces

Temperature
Viscosity Loss

sNonreversible
Rapture of Coil and
Subsequent Orientation
Under Shear Forces

Permanent
Viscosity Loss

FIG. 20Impact of shear stress on VI improver molecule.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER 13: HYDRAULIC FLUIDS 373


TABLE 9Low temperature viscosity
grades for hydraulic fluid classifications.
Viscosity
Grade

Min.

Max.

L5
L7
LIO
L15
L22
L32
L46
L68

-49
-41
-32
-22
-14
-7
-1

-50
-42
-33
-23
-15
-8
-2
4

that will crystallize at low temperatures and cause the fluid to


solidify. These additives do not entirely prevent wax from
crystallizing in the oil. Rather, they lower the temperature at
which large wax crystal structures are formed. By reducing
the size of the crystal matrix, pour point depressants permit
lubricants to flow at lower temperatures.
Two widely used types of pour point depressants are alkylaromatic polymers, which adsorb on wax crystals to inhibit
growth and adherence of crystals to each other, and polymethacrylates (PMA), which co-crystallize with wax to minimize growth of crystals. Depending mainly on the type of
base fluid, the pour point of oil can be lowered typically by
20-30F (I1-17C).

4684)all of which include specified cooling cycle and low


shear rates simulating field conditionscan be used to assess
fluid pumpability. Performance specifications that include
low temperature pumpability requirements, such as ASTM D
6080, Standard Practice for Defining the Viscosity Characteristics of Hydraulic Fluids, typically specify a temperature
range for different viscosity grades. In Table 9 (from Standard D 6080) the temperature range for a given L-grade is approximately equivalent to that for an ISO grade of the same
numerical designation and having a viscosity index of 100.
For instance, the temperature range for the L32 oil is approximately the same as an ISO VG 32 grade with a Viscosity
Index of 100.
Pour Point

Depressants

Pour point depressants are additives designed to reduce formation of rigid wax crystals in the lubricant at low temperatures. Conventionally refined mineral oils typically require
the use of pour point depressants because they contain wax

TYPES OF HYDRAULIC FLUIDS


The major compositional categories of hydraulic Quids are
Petroleum Based, Synthetic Based, Water Based, Vegetable
Oil Based, and Water (Fig. 21). As expected, these different
categories have properties that make them especially desirable in particular applications. In this section, the types of
hydraulic fluids will be discussed in terms of their defining
functionality rather t h a n composition. For example, fire resistant fluids, which are typically water-based or ester-based
fluids having high flash points, are used extensively in the basic metals industry where the risk of ignition is high, while
"environmentally acceptable" fluids are used in environmentally sensitive areas.
Hydraulic system hardware is usually designed and formally rated to work with mineral oils, since they are the
predominant hydraulic fluid in use. Systems may have to

HVDRMjuc mxs

Pelrolaum 8al

R&OlfWbHed

SynlHtBc Ba9Cl

AnCweartAW)

MuBgrade

Witer Baaed

SynttieBc Hydiocarfcona (SHC)

PoW/'}

Rapwwed

Polyalphaolat>i8(PAO)

ScBean

wsot 0 ^

Polyaloxarwa

EMere

Phosphale Eaan

SMeones

PololEgre

EWare

HatoaanalKl Cemiiounds

Polyetherg

Aryieftera

SynHiellc SoluUon

Water-h-QI (hvBrt)

O U n - W W f ;So>i>le OB

Ghlorinatad Hydrocarbons
Raflutor

Silanas

FluowEstere

V 9 a * f c b 0 1 Based

WiiBd Haloflan Coropounds

FIG. 21Schematic of hydraulic fluid types.


Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

Mcra Emulalm

374

MANUAL

37: FUELS AND LUBRICANTS

HANDBOOK

be modified to accominodate alternative types of hydraulic


fluids.

precipitates that can cause hydraulic valve sticking and filter


plugging [65].
HM

Mineral Hydraulic Oils


The majority of hydraulic fluids in service are mineral oil
based because they generally provide excellent performance
at a relatively low cost. Within the mineral hydraulic oil category there is a wide range of viscosity grades and fluid types.
The International Organization for Standardization (ISO) established a classification system for hydraulic fluids that is
designated ISO 6743-4: 1999, Lubricants, Industrial Oils and
Related Products (Class L)-Classification-Part 4: Family H
(Hydraulic Systems). Using this classification system as a
foundation, ASTM created ASTM D 6158, Standard Specification for Mineral Hydraulic Oils, which defines the physical
properties and performance r e q u i r e m e n t s of mineral hydraulic fluids. Table 10 provides a list of mineral oil based
fluids listed in ASTM D6158.
HH
Type H H fluids are straight base oils without any additives.
They may be used in air-over-oil hydraulic systems such as is
found in car lifts at automotive service centers. They are also
used in manual hydraulic pumps, jacks, and other low-pressure hydraulic systems. While type H H fluids are able to perform the primary function of a hydraulic fluid, which is to
transmit power, they are unable to withstand high temperatures and have limited lubricating capabilities. Thus these
fluids find limited application in industry.
HL
Type HL fluids are also mineral oil based, but they contain
rust and oxidation inhibitors to protect equipment from the
detrimental effects of water contamination and chemical deterioration due to heat. These fluids are also known as R & O
oils because they contain rust and oxidation inhibitors. Type
HL fluids are often recommended for use in machine tool applications where system pressures are limited to 2000 psi or
less. They are also recommended for some piston p u m p applications. For example, type HL fluids are the preferred fluid
for Denison piston p u m p s [63]. This is because some ZDTP
containing oils can be aggressive to yellow metal (brass and
bronze) and silver alloyed components in piston pumps.
R & O oils often are formulated using a rust inhibitor
chemistry that contains succinic acid derivatives [64]. These
additives may be incompatible with metallic sulfonate or
phenate rust inhibitors or ZDTP antiwear additives used in
many antiwear hydraulic fluids, resulting in formation of

TABLE 10Mineral oil based hydraulic fluid classifications.


Symbol

Classification

HH
HL

Non-inhibited refined mineral oils


Refined mineral oils with improved
rust protection and oxidation
stability
Oils of the HL type with improved
anti-wear properties
Oils of the HM type with improved
viscosity index properties

HM
HV

Commercial
Designation
Straight base oils
R&O oils
Antiwear oils
Multigrade oils

Type HM fluids contain antiwear additives in addition to the


rust and oxidation inhibitors found in HL fluids. They are the
most widely used mineral oil based hydraulic fluids because
antiwear additives provide enhanced performance in highpressure hydraulic applications. The requirements for HM
oils are listed in Table 11. While early versions of HM oils
lacked the thermal stability necessary for satisfactory piston
p u m p performance, modern fluids are able to perform quite
well in piston p u m p applications.
Zinc dialkyldithiophosphate is the most widely used antiwear additive for hydraulic applications. In recent years, concerns about the environmental effects of ZDTP have led to development of zinc-free or ashless antiwear hydraulic fluids.
These products utilize sulfur and phosphorus compounds to
achieve satisfactory antiwear performance. Thus, a type HM
fluid may contain zinc or some other type of antiwear additive chemistry.
HV
Type HV fluids contain the same basic chemistry as HM fluids plus a viscosity index (VI) improver. (See the Shear Stability section.) Viscosity index improvers impart multigrade
functionality to type HV fluids. While a wide range of polymers may be used for VI enhancement, these additives all
function in the same basic manner. At low temperatures, VI
improvers have a minimal effect upon fluid viscosity and at
high temperatures they have a thickening effect. This enables
the fluid to provide satisfactory performance at a wider operating temperature range [66].
T r a c t o r F l u i d s , ATF, a n d E n g i n e O i l s
Tractor fluids are unique in that they are formulated to lubricate transmissions, final drives, wet brakes, clutches, and
hydraulic systems from a c o m m o n fluid reservoir on the tractor [67]. Consequently, tractor fluids are often used in agricultural equipment, off-highway machinery, backhoes, and
turf applications where a multifunctional hydraulic fluid is
required. To lubricate gears, wet brakes, and hydraulic systems, tractor fluids typically utilize phosphate ester based
friction modifiers and ZDTP.
Automatic transmission fluids (ATFs) are similar to tractor
fluids in that they are designed for multiple functionality,
however, they generally utilize ashless antiwear additives
r a t h e r t h a n ZDTP. ATFs tend to be used in applications
where superior low temperature performance is desired because they are designed to remain fluid at temperatures as
low as 40C. Years ago it was common to use lOW engine
oils in hydraulic applications. Until recently lOW diesel engine oil was the primary hydraulic fluid recommendation for
Caterpillar equipment because lOW diesel engine oil contains ZDTP antiwear additives and is compatible with engine
oil [68]. The disadvantage of using ATFs and engine oils in
hydraulic applications is that they tend to have poor water
separation properties, which reduces wet filterability performance due to hydrolysis of the metallic sulfonates and phenates. Consequentiy, fluids designed specifically for hydraulic
performance are generally more desirable.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER

13: HYDRAULIC

FLUIDS

375

TABLE 11International specifications for antiwear hydraulic oil.


DIN 51524 Part 2

ISO 11158
ASTM
Properties
ISOVG
Kinematic Viscosity, cSt
D445
40'C max.
40C min.
D92
Flash Point, C min.
Flash Point, C min.
093
Brookfield Vis < 750 cP, Max
D2983
D97
Pour Point, C, max
Visual
Appearance
D1744
Water Content, wt%
01298
Density @1SC
TAN, ma KOH/g, max
0664 or 0974
D665A
Rust Test A
Rust Test B (Salt Water)
06658
Cu Rating (3 hr, 100C), max. 0130
TOST Oxidation, hours
0943
AN after 1000h, max.
Air Release @ 50C, minutes
D3427
(max)
Foam tendency/stability
0892
Seql
max
max
SeqII
max
SeqIII
Demulsibility @ 54C
01401
Minutes to 37 mL water .
0 5182
FZG Fail Stage
Ctiange in Hardness
0 471
NBR1,168 hrs @ 100C
ISO 7619
Change in Volume (%)
NBR1,168hrs@100C
Viscosity Index, min
0 2270
CM Thermal Stab.
02070A
Sludge, mg/100 ml max
Copper rod appearwice, rating (max.)
0 4310
Oxidation (1000 h)
AN, mgKOH/g max
Total sludge, mg max.
Total Metals in oil/water/sludge
Vickers 104C mg. wear max.
0 2882

HM (Antiwear)
32
46

22

S8

22

68

24.2
19.8
165

35.2
28.8
175

50.6
41.4
185

74.8
61.2
195

24.2
19.8
165

35.2
28.8
175

50.6
41.4
185

74.8
61.2
195

-18

-15
< Report >

-12

-12

-21

-18

-15

-12

-15
-21

-8
-18
<C&B>

-2
-15

4
-12

2
1000

< Report >


< Report >
<Pass>
< Pass>
2
1000

2
1000

2
1000

10

13

30

30

Oto-7

Oto-6

<-ao5->

< Report >


< Report >
< Report >
<Pass>

< Report >


< Report >
<Pass>

10

13

10

10

<15Qro>
<7SI0>

<i5ao>
<75/Q>
<150)>
30
10

<15C/0>
<70>
<150)>

<i5oro>
30
10

30
10

30
10

40

40
10

40
10

60
10

< Report >

Oto-8

Oto-7

Oto-7

Oto-6

< Report >


< Report >

0to15

0to12

0to12

OtolO

2
< Report >

HFAE

Oil-in-water emulsions containing


typically >80% water
Chemical solutions in water
containing typically >80% water
Water-in-oil emulsions containing
approximately 45% water
Water-polymer solutions containing
approximately 45% of water
Synthetic fluids containing no water
and consisting of phosphate esters
Synthetic fluids containing no water
and of other compositions

< 150 in exterxled test>

Commercial
Descriptions

Soluble oils
High water based
fluids
Invert emulsions
Water-glycols
Phosphate esters
Polyol esters

30

30

r288 hrs. f 000)


Oto-8
Oto-7
Oto15

0to12
<-90->

0to12

OtolO

< Report >


< Report >

25
5

25
5

25
5

200

200
< Report >
50

200

200

50

50

: Report >

Classification

HFDU

HM (Antiwear)
32
46

74.8
61.2
180
168

Symbol

HFDR

22

50.6
41.4
180
168

TABLE 12ISO designations for fire resistant


hydraulic fluids.

HFC

68

35.2
28.8
160
148

Fire resistant hydraulic fluids are used in the basic metals industry, die casting, military, and foundry applications. They
may be found in any application where a ruptured hydraulic
line presents a potential fire hazard. Fire resistant hydraulic
fluids are formulated with materials that have a lower BTU
content than mineral oils, such as polyol esters, phosphate esters, and water-glycol solutions. As a result, they b u m with
less heat generation than mineral hydraulic oils. As with mineral hydraulic fluids, the International Organization for Standardization has established a classification system for fire resistant fluids based upon composition. Table 12 provides a list

HFB

ASTM D6158

24.2
19.8
140
128

Fire Resistant Fluids

HFAS

Requirements
32
46

50

of the ISO designations for fire resistant hydraulic fluids [69].


While power transmission, heat transfer, and lubrication
are essential requirements for all types of hydraulic fluids, it
is sometimes necessary to compromise these properties to
accommodate a critical fluid characteristic. This is especially
true of fire resistant hydraulic fluids. Fire resistant fluids differ from mineral hydraulic fluids in density, compatibility,
and lubricating properties. As a result, hydraulic systems are
often modified when utilizing a fire resistant fluid. To optimize the performance of fire resistant fluids, the National
Fluid Power Association and ISO have published guides for
their use [70,71]. These NFFA and ISO documents detail the
operational characteristics of fire resistant fluids and provide
suggestions for storage, use, and handling of these fluids.
Table 13 provides a comparison of the properties of common
fire resistant hydraulic fluids.
HFA
HFA fluids contain greater than 80% water. These products
are sometimes referred to as 95:5 fluids, because 5% concentrations are commonly employed. The ISO 6743-4 classification divides HFA into two sub-categories: HFAE and HFAS.
HFAE fluids are oil-in-water emulsions. HFAS fluids are
chemical solutions or blends of selected additives in water.
Typically these products are sold as concentrates and diluted
prior to use in service. Because of the high vapor pressure of
water, the m a x i m u m recommended bulk fluid temperature
for HFA fluids is 50C [72]. The antiwear properties of these
fluids are inferior to mineral hydraulic fluids because the vis-

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

376

MANUAL

37: FUELS AND LUBRICANTS

HANDBOOK

TABLE 13Comparison of c o m m o n fire resistant fluid properties.


Property
ISO Designation
Heat of Combustion"
Autoignition Temp, "F*
Maximum" Temperature
Vapor Pressure, m b a r
Specific Gravity
Viscosity @ 40C, cSt
Water
Content
Vane p u m p rating"
Compatible Seals

Antiwear Hyd.
Oil

Invert
Emulsion

Water
Glycol

Phosphate
Ester

Polyol
Ester

HM
29.1 kJ/g
650
150F
0.001 @ 50C
0.85-0.88
32-68
0.05%

HF-B
16.3 kJ/g
830
120F
NA
0.91-0.93
80-100
43%

HF-C
5.3 kJ/g
830
120F
80 @ 50C
1.05-1.10
40
43%

HF-DR
19.0 kJ/g
1100-H
150F
< 1 @ 150C
1.02-1.16
22-100
0.05%

HF-DU
21.1 kJ/g
750"
150F
NA
0.91-0.96
46-68
0.1%

100%
Buna-N, Viton

33%
Nitroxyl, Buna-N

67%
Buna-N

67%
Butyl, EPR

100%
Viton, Buna-N

"Roberts and Brooks Flammability Data, NFPA T2.13.8-1997, a calculated estimate was used for HFDU.

cosity of HFA fluids is comparable to water, approximately 1


cSt. Performance is satisfactory with HFA fluids when suitable components are used but is apt to be poor if used in conventional hydraulic systems. Special precautions also are required in the selection of filter construction materials and
plumbing of p u m p inlets. Thus, it is necessary to work closely
with fluid a n d c o m p o n e n t suppliers w h e n utilizing HFA
fluids.
HFB
HFB fluids are water-in-oil emulsions consisting of
p e t r o l e u m oil, emulsifiers, selected additives, and water.
They are commonly referred to as invert emulsions. In an invert emulsion the oil phase, which provides lubricity and rust
protection, encapsulates the water phase, which provides fire
resistance. The water content of an HFB fluid is normally in
the 4 3 - 4 5 % range (w/w). When water content of these fluids
drops below 38% due to evaporation, the fire resistance of the
invert-emulsion deteriorates. Maintenance of invert emulsions is complicated by the fact that when these fluids lose
water through evaporation, a high-shear mixing device is
normally necessary for proper addition of make-up. The viscosity properties of invert emulsions are u n u s u a l in that
evaporation of water results in a viscosity decrease.
Several ASTM methods have been developed specifically
for invert emulsion hydraulic fluids. ASTM D 3709, Standard
Test Method for Stability of Water-in-Oil Emulsions Under
Low to Ambient Temperature Cycling Conditions, is used to
evaluate the freeze-thaw stability of invert emulsions. ASTM
D 3707, Standard Test Method for Storage Stability of Waterin-Oil Emulsions by the Oven Test Method is used to determine if the emulsion has a propensity to separate after 48 h
at 85C. As with HFA fluids, special precautions also are required in the selection of filter construction materials and
plumbing of p u m p inlets. Thus, it is necessary to work closely
with fluid and c o m p o n e n t suppliers w h e n utilizing HFB
fluids.
HFC
HFC Fluids are solutions of water, glycols, additives, and
thickening agents. They are commonly referred to as waterglycol hydraulic fluids. Typically, water-glycol fluids are formulated with diethylene glycol or propylene glycol and a
polyalkylene glycol based thickening agent [73]. The low
molecular weight glycol reduces the vapor pressure of tlje

fluid (relative to water) while high molecular weight


polyalkylene glycol acts as a thickening agent, much like a
viscosity index improver. This combination thickeners and
glycols enhance the lubricating properties of a water-glycol
and reduces the propensity of the fluid toward cavitation erosion. Nonetheless, operating temperatures for water-glycols
are limited to a maximum of 50C because of the effect of
temperature on vapor pressure [74].
Water glycol fluids are highly alkaline due to the presence
of amine based corrosion inhibitors. As a result, these fluids
can attack zinc, cadmium, magnesium, and non-anodized
aluminum, forming sticky or gummy residues. Consequently,
these metals should be avoided when selecting system components. Special precautions also are required in the selection of filter construction materials and plumbing of p u m p
inlets. Thus, it is necessary to work closely with fluid and
component suppliers when utilizing HFC fluids.
HFD
HFD Fluids are non-water containing fire resistant fluids.
The first edition of International Standard ISO 6743-4 classification (1982) divided HFD into four sub-categories:
HFDR, HFDS, HFDT, and HFDU. In 1999 the standard was
revised, deleting the HFDS and HFDT fluids from the classification system. HFDS and HFDT fluids are no longer
commercially viable because they were based upon chlorinated materials such as polychlorinated biphenyls (PCBs)
or other chlorinated aromatic compounds. Environmental
concerns associated with chlorinated hydrocarbons led
to withdrawal of these products from the market. On the
other hand, HFDR and HFDU fluids continue to be widely
used in a variety of commercial and military hydraulic applications.
HFDR fluids are composed of phosphate esters. The majority of phosphate ester type hydraulic fluids used in industrial applications are based upon triaryl phosphate [75].
Trialkyl and mixed alkylaryl phosphate esters are used in
aviation because of their lower density [76]. Phosphate esters are difficult to ignite because they are non-volatile and
chemically stable. The stability of p h o s p h a t e esters is
demonstrated by the fact that they do not propagate a flame
in the Standard Test Method for Linear Flame Propagation
Rate of Lubricating Oils and Hydraulic Fluids (ASTM D
5306-92). The principal reason they do not propagate a
flame is that the chemical reactions that take place during

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER
combustion of phosphate esters are endothermic. Thus,
phosphate esters generate less heat when burned relative to
other HFD fluids. In addition, because their Are resistance
is not dependent upon the presence of water or mist suppressing additives, the fire resistance of HFDR fluids does
not degrade in service.
HFDR fluids have been used in hydraulic applications for
more than forty years and are known for excellent inherent
lubricating properties [77]. In fact, aryl phosphate esters
serve as antiwear additives in mineral oil based hydraulic fluids [78]. However, phosphate esters have a steep viscosity
temperature curve, which makes their temperature operating
window rather narrow [79]. Hydrolysis is the most c o m m o n
form of degradation in HFDR fluid, and can occur in the
presence of a small amount of water and heat. When hydrolysis takes place, phosphate esters break down into their constituent acids and alcohols. Due to the frequent presence of
water in hydraulic applications, the sensitivity of phosphate
esters to water has limited their use and significantly reduced
their service life.
Phosphate esters are compatible with all common metals
except aluminum. Phosphate esters do not "wet" the surface
of aluminum and thus aluminum should not be used in tribological contacts such as bearings [80]. Phosphate esters
should never be added to systems containing mineral oil or
water-based fire resistant fluids. Not only are these materials
chemically incompatible with each other, in all probability
preexisting gaskets, seals, hoses, and coatings are also incompatible. Special precautions also are required in the selection of filter construction materials and plumbing of
p u m p inlets. Thus, it is necessary to work closely with fluid
and component suppliers when utilizing HFD fluids.
HFDU fluids typically are composed of polyol esters although other materials such as polyalkylene glycols are included in the HFDU category. Trimethylol propane oleate,
neopentyl glycol oleate, and pentaerythritol esters are the
most c o m m o n of the synthetic polyol esters. Triglycerides
derived from soybeans, sunflower, and rapeseed plants are
naturally occurring polyol esters that also are used in HFDU
fluids. Polyol esters derive their fire resistance from a combination of factors. First, polyol esters have a relatively high
flash, fire, and autoignition point. Second, they b u m with
less energy than oil because of the presence of oxygen in the
molecule. And finally, polyol ester fire resistant fluids employ antimist additives that enhance their spray-flammability resistance [81]. Depending upon the shear stability of the
polymer, the fire resistance of the fluid may deteriorate in
service.
Like phosphate esters, polyol esters have excellent lubricating properties but are prone to hydrolysis in the presence
of water [82]. In addition, they are vulnerable to oxidation because of unsaturation irl the fatty acid portion of the ester.
These factors tend to limit their service life relative to mineral
oils. Most common metals used in hydraulic applications are
compatible with polyol ester hydraulic fluids, with the exception of lead, zinc, and cadmium. Unlike other fire resistant
fluids, polyol esters performance is satisfactory with comm o n filter construction materials and system designs. Thus it
is relatively easy to convert a hydraulic system that operates
on mineral oil based hydraulic fluids to HFDU fluids.

13: HYDRAULIC

FLUIDS

377

Environmentally Acceptable Hydraulic Fluids


Environmentally acceptable hydraulic fluids have found
their way into hydraulic applications where there is risk of
fluid leaks and spills entering the environment (especially
waterways) affecting aquatic and terrestrial life. Some examples of these niche markets include forestry, construction,
locks a n d d a m s , heavy-duty lawn equipment, a m u s e m e n t
parks/entertainment industry, offshore drilling, and maritime. Most environmentally acceptable hydraulic fluids
exhibit two key environmental characteristics: virtual nontoxicity to aquatic life a n d aerobic biodegradability. Organizations such as the Organization for Economic Co-operation
and Development (OECD), the Co-ordinating E u r o p e a n
Council (CEC), and the U.S. Environmental Protection
Agency (EPA) have developed standard test methods to determine the toxicity a n d biodegradability of substances.
More recently ASTM has developed a Guide for Assessing
Biodegradability of Hydraulic Fluids (ASTM D 6006) and a
Classification of Hydraulic Fluids for Environmental Impact
(ASTM D 6046) based on the above organizations' methods.
Utilizing the methodology from these organizations, standard classifications and performance requirements for environmental fluids have also been established by the International Organization for Standardization (ISO) and regional
environmental organizations t h a t a w a r d Eco Labels (i.e.,
German Blue Angel, Nordic Swan, Japanese EcoMark). ISO
environmental hydraulic fluid classifications are described in
Table 14.
HETG
Type HETG fluids are based on naturally occurring vegetable
oils or triglyceride esters. Without the addition of a thickener, vegetable oils are limited to a narrow viscosity range between ISO 32 and 46. While HETG fluids biodegrade rapidly,
have excellent natureil lubricity and have a natural VI in excess of 200, they are unsuitable for use at high and low temperature extremes. This is because they tend to gel at low
temperatures and oxidize at high temperatures. The practical
temperature limits for uses HETG fluids is 25F to 165F.
HEES
Type H E E S fluids are based on unsaturated to fully saturated
synthetic esters. Common ester chemistries utilized for hydraulic fluids consist of TMP oleates, neopentylglycols, pentaerythritol esters, adipate esters, and complex esters. The
synthetic esters provide better performance over HETG t5T3e
hydraulic fluids with wider operating temperature ranges,
broad range of ISO viscosity grades, and better oxidation stability while still maintaining biodegradability.

TABLE 14ISO environmental hydraulic fluid classifications.


Symbol

Classification

Commercial Designation

HETG
HEES

Vegetable oil types


Sjmthetic ester types

HEPG
HEPR

Polyglycol types
Polyalphaolefln types

Vegetable oils and natural esters


Polyol esters, neopentylglycols,
syntiietic adipate esters
Polyglycols
Polyalphaolefins (PAO) or
synthetic hydrocarbons (SHC)

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

378 MANUAL 37: FUELS AND LUBRICANTS HANDBOOK


HEPG
Type HEPG fluids are polyethyleneglycols (PEG), which possess good oxidation stability and low temperature flow characteristics. At molecular weights of up to 600-800, HEPG
type fluids are ecotoxicologically harmless and readily
biodegradable (>90% in 21 days) [83]. Some disadvantages
of this class of fluids include miscibility with water, incompatibility with mineral oils, and aggressiveness toward some
common t5^es of elastomer seal materials.
HEPR
HEPR type fluids are polyalphaolefins (PAO) or synthesized
hydrocarbon (SHC) base fluids, which have significantly better viscometric properties over a wider range of temperatures
than mineral base fluids with the same standard viscosity
classification. Some low viscosity PAOs have shown acceptable primary biodegradability, though not as rapid as vegetable or synthetic ester base fluids (Fig. 22). Additional advantages claimed for synthetic lubricants over comparable
petroleum-based fluids include improved thermal and oxidative stability, superior volatility characteristics, and preferred
frictional properties.

Another challenge that comes with the various hydraulic


applications is that of developing test methods that are truly
representative of performance in actual systems. Bench-top
tests are to be used as logical indicators of a fluid's response
to expected conditions of temperature, pressure, contamination, etc. A significantly higher number of variables concurrently influence the fluid more than any single bench test can
simulate. Therefore, standards and specifications consist of
multiple bench tests as well as more realistic full-scale test
stands that use actual pumps in typical hydraulic circuits.
Test methods will continue to evolve as more sophisticated
techniques are developed to predict field performance of hydraulic fluids.

ASTM STANDARDS
No.
D 92
D 95
D 96
D 97
D 130

CONCLUSIONS
A well formulated hydraulic oil consists of a properly selected
base fluid and the appropriate balance of additives, optimized to provide the best possible overall performance required for the targeted application. The versatility of hydraulics makes fluid power advantageous in a wide variety of
industrial and mobile applications. With this versatility
comes the challenge of developing fluids that function appropriately in a wide range of conditions, even as environmental health and safety requirements become more and
more stringent. New fluid technologies continue to emerge to
meet these challenges.

D 287
D 445
D 446
D 471
D 664

Title
Test Method for Flash and Fire Points by Cleveland
Open Cup
Test Method for Water in Petroleum Products and
Bituminous Materials by Distillation
Test Method for Water and Sediment in Crude Oil
by Centrifuge Method
Test Method for Pour Point of Petroleum Products
Test Method for Determination of Copper Corrosion from Petroleum Products by the Copper Strip
Tarnish Test
Test Method for API Gravity of Crude Petroleum
and Petroleum Products (Hydrometer Method)
Test Method for Kinematic Viscosity of Transparent and Opaque Liquids (the Calculation of Dynamic Viscosity)
Specifications and Operating Instructions for Glass
Capillary Kinematic Viscometers
Test Method for Rubber Property-Effect of Liquids
Test Method for Acid Number of Petroleum Products by Potentiometric Titration

Polypropylene glycols

I Mininnum
Maximum

Mineral oils

Hydro-treated
mineral oils
Polyethylene glycols
1 Vegetable oils

Synthetic esters
20

40

60

80

100%

FIG. 22Chart comparing primary biodegradation of base fluids by CEC


method.
Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER
D 665
D 892
D 943
D 974
D 1298

D 1401
D 1744
D 2070
D 2270
D 2271
D 2272
D 2422
D 2619
D 2717
D 2766
D 2783
D 2882

D 2983

D 3339
D 3427
D 3519
D 3603

D 3707
D 3709

D 4172
D 4310
D 4684

D 5133

D 5182

Test Method of Rust-Preventing Characteristics of


Inhibited Mineral Oil in the Presence of Water
Test Method for Foaming Characteristics of Lubricating Oils
Test Method for Oxidation Characteristics of Inhibited Mineral Oils
Test Method for Acid and Base Number by ColorIndicator Titration
Test Method for Density, Relative Density (Specific
Gravity), or API Gravity of Crude Petroleum Products by Hydrometer Method
Test Method for Water Separability of Petroleum
Oils and Synthetic Fluids
Test Method for Determination of Water in Liquid
Petroleum Products by Karl Fischer Reagent
Test Method for Thermal Stability of Hydraulic Oils
Practice for Calculating Viscosity Index from Kinematic Viscosity at 40C and 100C
Test Method for Preliminary Examination of Hydraulic Fluids (Wear Test)
Test Method for Oxidation Stability of Steam Turbine Oils by Rotating B o m b
Classification of Industrial Fluid Lubricants by Viscosity System
Test Method for Hydrolytic Stability of Hydraulic
Fluids (Beverage Bottle Method)
Test Method for Thermal Conductivity of Liquids
Test Method for Specific Heat of Liquids and Solids
Test Method for Measurement of Extreme-Pressure
Properties of Lubricating Fluids (Four-Ball Method)
Test Method for Indicating the Wear Characteristics of Petroleum a n d Non-Petroleum Hydraulic
Fluids in a Constant Volume Vane Pump
Test Method for Low-Temperature Viscosity of Automotive Fluid Lubricants Measured by Brookfield
Viscometer
Test Method for Acid N u m b e r of Petroleum Products by Semi-Micro Color Indicator Titration
Test Method for Air Release Properties of Petroleum
Oils
Test Method for Foam in Aqueous Media (Blender
Test)
Test Method for Rust-Preventing Characteristics of
Steam Turbine Oils in the Presence of Water (Horizontal Disk Method)
Test Method for Storage StabiHty of Water-in-Oil
Emulsions by the Oven Test Method
Test Method for Stability of Water-in-Oil Emulsions
Under Low to Ambient Temperature Cycling Conditions
Test Method for Wear Preventive Characteristics of
Lubricating Fluid (Four-Ball Method)
Test method for Determination of the Sludging and
Corrosion Tendencies of Inhibited Mineral Oils
Test Method for Determination of Yield Stress and
Apparent Viscosity of Engine Oils at Low Temperatures
Test Method for Low temperature. Low Shear Rate,
Viscosity/Temperature Dependence of Lubricating
Oils Using a Temperature Scanning Technique
Test Method for Evaluating the Scuffing Load Ca-

D 5306
D 5534
D 5621
D 6006
D 6046
D 6080
D 6158
D 6278

D 6351
D 6546

D 6547

13: HYDRAULIC

FLUIDS

379

pacity of Oils (FZG Visual Method)


Standard Test Method for Linear Flame Propagation Rate of Lubricating Oils and Hydraulic Fluids
Test Method for Vapor-Phase Rust-Preventing
Characterisitics of Hydraulic Fluids
Test method for Sonic Shear Stability of Hydraulic
Fluid
Guide for Assessing Biodegradability of Hydraulic
Fluids
Classification of Hydraulic Fluids for Environmental Impact
Practice for Defining the Viscosity Characteristics
of Hydraulic Fluids
Specification for Mineral Hydraulic Oils
Test Methods for Shear Stability of Polymer Containing Fluids Using a European Diesel Injector Apparatus
Test Method for Determination of Low Temperature Fluidity and Appearance of Hydraulic Fluids
Test Methods for and Suggested Limits for Determining Compatibility of Elastomer Seals for Industrial Hydraulic Fluid Applications
Test Method for Corrosiveness of a Lubricating
Fluid to a Bi-Metallic Couple

OTHER STANDARDS
AFNOR NF E48-690: Hydraulic Fluid Power. Fluids. Measurement of Filtrability of Mineral Oils
AFNOR NF E48-691: Hydraulic Fluid Power. Fluids. Measurement of Filtrability of Minerals Oils in the Presence of
Water
ANSI/(NFPA) S t a n d a r d T2.13.7R1-1996: Hydraulic Fluid
Power - Petroleum Fluids - Prediction of Bulk Moduli
ISO 6743/4 Part 4: Family H (Hydraulic Systems), Lubricants, Industrial Oils and Related Products (Class L ) : Classification Part 4: Family H (Hydraulic Systems)
ISO 12922: Lubricants, Industrial Oils, and Related Products
(Class L)Family H (Hydraulic systems)Specifications
for categories HFAE, HFAS, HFB, HFC, HFDR and HFDU
ISO/DIS 15380: Lubricants, Industrial Fluids and Related
Procedures (Class L), Family H (Hydraulic Systems)-Specifications for Catagories HETG, HEPG, HEES and HEPR

REFERENCES
[1] Halliday, D. and Resnick, R. Physics, 3rd ed., John Wiley &
Sons, NY, 1978, p. 376.
[2] Henke, R., Diesel Progress, Fluid Power Buyer's Guide, Diesel
and Gas Turbine Publications, Waukesha, WI, 1994, p. 4.
[3] Vickers Industrial Hydraulics Manual, Ch. 3, Vickers, Inc.,
Rochester Hills, MI, 1992, p. 3-1.
[4] Booser, E. R., Handbook of Lubrication, Volume II Theory and
Design, CRC Press, Boca Raton, FL, 1987, pp. 242-244.
[5] NFPA Standard T2.13.7R1 - 1997: Hydraulic Fluid Power Petroleum Fluids - Prediction of Bulk Moduli, National Fluid
Power Association, Milwaukee, WI, 1997.
[6] Lightening Reference Handbook, 7th ed., Paul-Munroe Hydraulics, Whittier, CA, 1990, p.l22.
[7] Totten, G. E., Webster, G. M., and Yeaple, F. D., "Physical Prop-

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

380 MANUAL 37: FUELS AND LUBRICANTS HANDBOOK


erties and Their Determination," Handbook of Hydraulic Fluid
Technology, G. E. Totten, Ed., 2000, Marcel Dekker, NY, p. 253.
[8] Esposito, A., Fluid Power with Applications, Prentice Hall, Englewood Cliffs, NJ, 1988, p. 478.
[9] Michael, P. W. and Wanke, T. S, "Surgically Clean Hydraulic
Fluid - A Case Study," Proceedings of the 47th National Conference on Fluid Power, National Fluid Power Association, Milwaukee, WI, 1996, pp. 129-136.
[10] Esposito, A., Fluid Power with Applications, Prentice Hall, Englewood Cliffs, NJ, 1988, p.3.
[11] Ruble, L. R., "The Expanded Focus, Use, and Future of Water
Powered Rotary Actuators," Proceedings of the 48th National
Conference on Fluid Power, National Fluid Power Association,
Milwaukee, WI, 2000, pp. 567-574.
[12] Perrier, R., "Lubricant Basestocks," Speech, STLE Chicago
Lube School, Downers Grove, IL, 17 Mar. 1999.
[13] Engine Oil Licensing and Certification System, 14th ed.. Publication 1509, American Petroleum Institute, Washington DC,
1996.
[14] Wills, J. G., Lubrication Fundamentals, Mobil Oil Corporation,
Fairfax, VA, 1980, p. 30.
[15] Denisov, E., Handbook of Antioxidants, CRC Press, Boca Raton,
FL, 1995, p. 19.
[16] Lubricant and Fuel Additives, Advanced Lubrication Education
Program, Report 1525, May 1992.
[17] Rasberger, M., "Chemistry and Technology of Lubricants," Oxidative Degradation and Stabilisation (Sic) of Lubricants,
Blackie and Academic Professional, London, 1992, pp. 83-123.
[18] Igarashi, J., "Oxidative Degradation of Engine Oils," Japanese
Journal of Tribology, Vol. 35, No. 10, 1990. p. 1098.
[19] Saxena, D., Mookken, R. T., Srivastava, S. P., and Bhatnagar A.
K., "An Accelerated Aging Test for AW Oils," Lubrication Engineering, Vol. 49, No. 10, 1993, pp. 801-809.
[20] Approved Lubricants Manual, Cincinnati Machine P u b No.
lO-SP-95046, Cincinnati Machine, Cincinnati, OH, 1995,
pp. 2-40.
[21] Reyes-Gavilan, J. L. and Odorisio, P., "A Review of the Mechanisms of Action of Antioxidants, Metal Deactivators and Corrosion Inhibitors," NLGI Spokesman, Vol. 64, No. 11, Feb. 2001,
pp. 22-33.
[22] Gatto, V. J. and Grina, M. A., "Effects of Base Oil Type, Oxidation
Test Conditions and Phenolic Structure on the Detection and
Magnitude of Hindered Phenol/Diphenylamine Synergism," Lubrication Engineering, Vol. 55, No. 1, Jan. 99, pp. 11-20.
[23] Pike, R. and Conway-Jones, J. M., "Friction and Wear of Sliding
Bearings," ASM Handbook, Volume 18, Friction, Lubrication,
and Wear Technology, 1992, ASM International, Materials Park,
OH, p. 519.
[24] Blau, P. J., "Glossary pf Terms," ASM Handbook, Friction, Lubrication, and Wear Technology, Volume 18, 1992, ASM International, Materials Park, OH, p. 11.
[25] Michael, P. W., Herzog, S. N., and Marougy, T. E., "Fluid Viscosity Selection Criteria for Hydraulic Pumps and Motors" Proceedings of the 48th National Conference on Fluid Power,
Chicago, National Fluid Power Association, Milwaukee, WI,
2000, pp. 313-325.
[26] ISO 3448: Industrial Liquid Lubricants-ISO Viscosity Classification, International Organization for Standardization, Geneva,
1992.
[27] Michael, P.W., "Lubrication Basics," Plant Services, Vol. 14, No.
5, May 1993, pp. 19-22.
[28] Silva, G., "Wear Generation in Hydraulic Pumps," presented at
the SAE International Off Highway Congress, Milwaukee, WI,
1990, Paper 901679, Society for Automotive Engineers, Warrendale, PA.
[29] Voitik, R. M., "Realizing Bench Test Solutions to Field Tribology Problems Utilizing Tribological Aspect Numbers," Tribol-

ogy - Wear Test Selection for Design and Application, ASTM


STP 1199, A. W. Ruff and R. G. Bayers, Eds., ASTM International, West Conshohocken, PA, 1993, pp. 45-59.
[30] Hogmak, S. and Jacobson, S., "Hints and Guidelines for Tribotesting and Evaluation," Lubrication Engineering, Vol. 48,
No. 5, May 1992 p. 401.
[31] ASTM Round Robin Test Program RR: D02-1152, ASTM International, West Conshohocken, PA, 1994.
[32] Xie, L., Bishop, Jr., R. J., and Totten, G. E., "Bench and P u m p
Testing of Hydraulic Fluids," H a n d b o o k of Hydraulic Fluid
Technology, G. E. Totten, Ed., Marcel Dekker, NY, 2000, p. 526.
[33] Perez, J. M., Klaus, E. E., and Hansen, R. C , "Comparative Evaluation of Several Hydraulic Fluids in Operational Equipment, A
Full-scale Pump Stand Test and the Four-Ball Wear Tester. Part
II - Phosphate Esters, Glycols and Mineral Oils," Lubrication
Engineering, Vol. 46, No. 4, April 1990, p. 249.
[34] Perez, J. M., Weller, D. E., and Duda, J. L., "Sequential Four-Ball
Study of Some Lubricating Oils," Lubrication Engineering, Vol.
55, No. 9, 1994, pp. 28-32.
[35] Reichel, J., "Importance of Mechanical Testing of Hydraulic
Fluids," Tribology of HydrauUc P u m p Testing, ASTM STP 1310,
George E. Totten, Gary H. Kling, and Donald J. Smolenski, Eds.,
ASTM International, West Conshohocken, PA, 1996.
[36] Xie, L., Bishop, Jr., R. J., and Totten, G. E., "Bench and P u m p
Testing of Hydraulic Fluids," H a n d b o o k of Hydraulic Fluid
Technology, G. E. Totten, Ed., Marcel Dekker, NY, 2000, p. 526.
[37] Denison Test E q u i p e m e n t and Instructions, Consigne TP30283, Denison Hydraulics, Marysville, OH, 1999.
[38] Kling, G. H., Totten, G. E., and Ashraf, A., "Strategies for Developing Performance Standards for Alternative Hydraulic Fluids,"
in Lubricants for Off-Highway Applications, SAE SP-1553, Society for Automotive Engineers, Warrendale, PA, 2000, pp. 1-9.
[39] Ohkawa, S., Konishi, A., Hatano, H., Ishihama, K., Tanaka, K.,
and Iwamura, M. "Oxidation and Corrosion Characteristics of
Vegetable-Base Biodegradable Hydraulic Oil," SAE Technical
Paper Series, Paper N u m b e r 951038, Society for Automotive
Engineers, Warrendale, PA, 1995.
[40] Melief, H. M., Totten, G. E., and Bishop, R. J., "Overview of the
Proposed Rexroth High-Pressure Piston P u m p Testing Procedure for Hydraulic Fluid Qualification," Fluid Power for OffHighway Applications, SAE SP-1380, Society for Automotive
Engineers, Warrendale, PA, 1998.
[41] "Sundstrand Water Stability Test Procedure," Sundstrand Bulletin No. 9658, Sauer-Danfoss Corp., Ames, lA.
[42] Hodges, P. K.B., Hydraulic Fluids, John Wiley, NY, 1996, p. 116.
[43] Karl Fischer Applications, Mettler-Toledo AG., CH-8606
Griefensee, Switzerland, 1992.
[44] Lloyd, B. J., "Water Water Everywhere," Lubes-N-Greases,
Vol.11, No. 3, 1997, pp. 44-53.
[45] Tsong-Dsu, L. and. Mansfield, J., "Effect of Contamination on
the Water Separability of Steam Turbine Oils," Lubrication Engineering, Vol. 51, No. 1, 1995, pp. 81-85.
[46] Barber, A. R. and Perez, R. J., Air Release Properties of Hydraulic Fluids, The Lubrizol Corporation, Wickliffe, OH.
[47] Totten, G. E., Sun, Y. H., Bishop, R. J., and Lin, X., "Hydraulic
System Cavitation: A Review," SAE 982036, Society for Automotive Engineers, Warrendale, PA, 1998.
[48] Brew, D., Khonsari, M., and Ball, J. H.,"Current Research in
Cavitating Fluid Films," STLE Special Publication SP-28, Society of Tribologists and Lubrication Engineers, Park Ridge, IL,
1990, pp. 63-65.
[49] "Foaming and Air Entrainment in Lubrication and Hydraulic
Systems," Mobil Technical Bulletin, Mobil Oil, Fairfax, VA,
1989.
[50] HVLP HydrauUc Oils, Minimum Requirements, DIN 51524 Part
II, Deutsche Norm, Beuth Verlag GmbH, BerUn, 1990.
[51] Denison Fluid Standard HF-0: Hydraulic Fluid-For Use in Axial

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

CHAPTER 13: HYDRAULIC FLUIDS 381

[52]
[53]
[54]

[55]

[56]
[57]

[58]
[59]

[60]

[61]

[62]

[63]
[64]

[65]

[66]

Piston Pumps and Vane Pumps in Severe Duty Applications,


Denison Hydraulics, Marysville, OH, 1995.
Claxon, P. D., "Aeration of Petroleum Based Steam Turbine
Oils," Tribology, Februaiy 1972, pp. 8-13.
Hatton, D. R., "Some Practical Aspects of Turbine Lubrication,"
The Canadian Lubrication Journal, Vol. 4, No. 1, 1984, p. 4.
Rhee, I., Velz, C , and Von Bemewitz, K., Evaluation of Environmentally Acceptable Hydraulic Fluids, Technical Report No.
13640, U.S. Army Tank-Automotive Command, Research Development and Engineering Center, Warren, MI, March 1995.
Rizvi, S. Q. A., "Lubricant Additives and Their Functions," ASM
Handbook 10th ed., Vol. 18, Friction, Lubrication, and Wear
Technology, ASM International, Materials Park, OH, 1992, pp.
98-112.
Available from ASTM International Headquarters, 100 B a r r
Harbor Drive, PO Box C700, West Conshohocken, PA.
F a i n m a n , M. Z. and Hiltner, L. G., "Compatibility of Elastomeric Seals and Fluids in Hydraulic Systems," Lubrication
Engineering, Vol. 37, May 1980, pp. 132-137.
Ashby, D. M., "O-ring Specifications - Some Important Considerations," Hydraulics & Pneumatics, August 1999, pp. 53, 54, 90.
Leslie, R. L. Sculthorpe, H. J., "Hydraulic Fluids Compatible
with Metalworking Fluids," Lubrication Engineering, Vol. 28,
May 1970, pp. 165-167.
Carpsjo, C , Proceedings of International Symposium on Performance Testing of Hydraulic Fluids, Institute of Petroleum,
London, 3-6 Oct 1978.
Stambaugh, R. L. a n d K o p k o , R. J. "Behavior of Non-Newtonian
Lubricants in High Shear Rate Applications," SAE Transactions
82, Paper 730487, Society of Automotive Engineers, Warrendale, PA, 1973.
Stambaugh, R. L., Kopko, R. J., and Roland, T. F., "Hydraulic
P u m p Performance - A Basis for Fluid Viscosity Classification,"
SAE International Off Highway Congress, Milwaukee, WI, Paper
901633, Society of Automotive Engineers, Warrendale, PA, 1990.
Hydraulic Fluids Information, Denison Corp., Marysville, OH,
1995, p. 2.
Rizvi, S. Q. A., "Lubricant Additives and Their Functions," ASM
Handbook Volume 18 Friction, Lubrication, and Wear Technology, ASM International, Materials Park, OH, 1992, p. 106.
Sharma, S. K., Snyder Jr., C. E., Gschwender, L. J., Lang J. C ,
and Schreiber, B. F., "Stuck Servovalves in Aircraft Hydraulic
Systems," Lubrication Engineering, Vol. 55, No. 7, July 1999.
Michael, P. W. and Webb, S., "Future Fluids - What's Coming
Down the Hydraulic Line," OEM Off-Highway, January 1998,
pp. 34-36.

[67] Lubrizol Ready Reference for Lubricant and Fuel Performance,


The Lubrizol Corporation, Wickliffe, OH, 1998, p. 111.
[68] Colver, R., Chek-Chart Publications, Simon & Schuster, NY,
1995, p. 45.
[69] ISO 6743-4: Lubricants, Industrial Oils and Related
Products (Class L) - Classification - Part 4: Family H (hydraulic
systems). International Organization for Standardization,
Geneva, 1999.
[70] NFPA/T2.13.1: R3-1997, R e c o m m e n d e d Practice - Hydraulic
Fluid Power - Use of Fire Resistant Fluids In Industrial Systems,
NFPA, Milwaukee, WI, 1997.
[71] ISO 7745: Hydraulic Fluid Power - Fire Resistant (FR) Fluids Guidelines for Use, International Organization for Standardization, Geneva, 1999.
[72] Vickers Guide to Alternative Fluids, Eaton Corp., Southfield,
MI, November 1992.
[73] Totten, G. E. and Sun, Y., "Water-Glycol Hydraulic Fluids,"
Handbook of Hydraulic Fluid Technology, G. E. Totten, Ed.,
Marcel Dekker, NY, 2000, pp. 917-982.
[74] NFPA/T2.13.1: R3-1997, Recommended Practice - Hydraulic
Fluid Power - Use of Fire Resistant Fluids in Industrial Systems,
NFPA, Milwaukee, WI, 1997.
[75] Phillips, W. D., "Phosphate Ester Hydraulic Fluids," in Handbook of Hydraulic Fluid Technology G. E. Totten, Ed., Marcel
Dekker, NY, 2000, pp. 1025-1027.
[76] Parker O-Ring H a n d b o o k Y2000 Edition, Parker Hannifin
Corp., Cleveland, OH, 1999, pp. 3-18.
[77] O'Connor, J. and Boyd, J., Standard Handbook of Lubrication
Engineering, McGraw-Hill, NY, 1968, pp. 2-16.
[78] MIL-H-5606G Military Specification: Hydraulic Fluid,
Petroleum Base, Aircraft, Missile and Ordinance, NATO code
n u m b e r H-515, U.S. Department of Defense, Fort Belvoir, VA,
1994.
[79] Parker O-Ring H a n d b o o k Y2000 Edition, Parker Hannifin
Corp., Cleveland, OH, 1999, pp. 3-18.
[80] Phillips, W. D., "Phosphate Ester Hydraulic Fluids," Handbook
of Hydraulic Fluid Technology, G. E. Totten, Ed., Marcel
Dekker, NY, 2000, pp. 1025-1027.
[81] Gere, R. A. and Hazelton, R. A., "Polyol Ester Fluids," Handbook
of Hydraulic Fluid Technology, G. E. Totten, Ed., Marcel
Dekker, NY, 2000, pp. 983-1022.
[82] Hohn, B. R., Michaelis, K., and Dobereiner, R., "Load Cjirrying
Capacity of Fast Biodegradable Gear Lubricants," Lubrication
Engineering, 1999, Vol. 55, p. 37.
[83] Bartz, W. J., Environmentally Acceptable Hydraulic Fluids,
Technische Akademie Esslingen, Ostfildern, Germany.

Copyright by ASTM Int'l (all rights reserved); Thu Jun 12 01:39:47 EDT 2014
Downloaded/printed by
UNIVERSITY TEKNOLOGI MALAYSIA pursuant to License Agreement. No further reproductions authorized.

You might also like