You are on page 1of 208

SELECTED TECHNICAL PAPERS

STP1573

Editor: John Sherman

Fire Resistant Fluids

ASTM Stock #STP1573

ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19438-2959.
Printed in the U.S.A.
Library of Congress Cataloging-in-Publication Data
Fire resistant fluids / editor, John Sherman.
pages cm
“ASTM Stock #STP1573.”
Includes bibliographical references and index.
ISBN 978-0-8031-7591-4 (alk. paper)
1. Fire testing. 2. Hydraulic fluids--Testing. 3. Fire resistant materials. I. Sherman, John. II. American
Society for Testing and Materials.
TH9446.H9F575 2014
628.9’22--dc23 2014033297

Copyright © 2014 ASTM INTERNATIONAL, West Conshohocken, PA. All rights reserved. This material
may not be reproduced or copied, in whole or in part, in any printed, mechanical, electronic, film, or
other distribution and storage media, without the written consent of the publisher.

Photocopy Rights
Authorization to photocopy items for internal, personal, or educational classroom use, or the internal,
personal, or educational classroom use of specific clients, is granted by ASTM International provided
that the appropriate fee is paid to the Copyright Clearance Center, 222 Rosewood Drive, Danvers,
MA 01923, Tel: (978) 646-2600; http://www.copyright.com/
The Society is not responsible, as a body, for the statements and opinions expressed in this publication.
ASTM International does not endorse any products represented in this publication.

Peer Review Policy


Each paper published in this volume was evaluated by two peer reviewers and at least one editor. The
authors addressed all of the reviewers’ comments to the satisfaction of both the technical editor(s) and
the ASTM International Committee on Publications.

The quality of the papers in this publication reflects not only the obvious efforts of the authors and the
technical editor(s), but also the work of the peer reviewers. In keeping with long-standing publication
practices, ASTM International maintains the anonymity of the peer reviewers. The ASTM International
Committee on Publications acknowledges with appreciation their dedication and contribution of time
and effort on behalf of ASTM International.

Citation of Papers
When citing papers from this publication, the appropriate citation includes the paper authors, “paper title”,
STP title, STP number, book editor(s), page range, Paper doi, ASTM International, West Conshohocken,
PA, year listed in the footnote of the paper. A citation is provided on page one of each paper.

Printed in Bay Shore, NY


September, 2014
Foreword
This Compilation of Selected Technical Papers, STP1573 on Fire Resistant Fluids,
contains nine papers presented at a symposium with the same name held in Montreal,
Quebec, Canada, on June 24, 2013 and JAI102180, published June, 2009, Volume 6,
Issue 6 and determined to be pertinent to the topic. The symposium was sponsored
by the ASTM International Committee D02 on Petroleum Products, Liquid Fuels,
and Lubricants and Subcommittee D02.N0.06 on Fire Resistant Fluids.
The Symposium Co-Chairpersons and STP Editors are John Sherman, BASF
Corporation, Canton, MI, USA and Betsy Butke, Lubrizol Corporation, Wickliffe,
OH, USA.
Contents

Overview vii

Use of Thickened High Water Hydraulic Fluid in Flat Rolled Steel Production 1
J. Sherman, J. Maloy, E. Martino, P. Cusatis, and P. Fasano

Fire Resistant Fuel for Military Compression Ignition Engines 24


S. R. Westbrook, B. R. Wright, S. D. Marty, and J. Schmitigal

Ion Exchange and Mechanical Purification of Fire-Resistant Phosphate Ester Fluids


Used in Steam-Turbine Control Systems 38
W. D. Phillips, J. W. G. Staniewski, and S. Suryanarayan

Phosphate Ester-based Fluid Specific Resistance: Effects of Outside


Contamination and Improvement Using Novel Media 75
M. G. Hobbs and P. T. Dufresne Jr.

Thirty-Seven Years of Fleet Operating and Maintenance Experience Using


Phosphate Ester Fluids for Bearing Lubrication in Gas-Turbine/Turbo-Compressor
Applications 93
P. T. Dufresne

Property and Performance Evaluation of Water Glycol Hydraulic Fluids 109


P. Cusatis, J. Sherman, P. Fasano, and R. Bishop

Anhydrous Fire-Resistant Hydraulic Fluids Using Polyalkylene Glycols 126


M. R. Greaves and A. Larson

Polyalkylene Glycol Hydraulic Fluids, 20 Years of Fire Resistance 143


K. P. Kovanda and M. Latunski

Performance Comparison of Non-Aqueous Fire-Resistant Hydraulic Fluids 155


S. Rea and D. Barker

Assessing and Classifying the Fire-Resistance of Industrial Hydraulic Fluids:


The Way Ahead? 181
W. D. Phillips
Overview
Fire-resistant fluids are an integral component to the safe operation of key processes
for many industries where fire is a major hazard. Industries using fire-resistant fluids
today include steel manufacturing, aluminum die-casting, automobile manufac-
turing, food processing and electrical and nuclear power utilities.
The requirements for fire-resistant fluids are not static but dynamic as industries
using these fluids in turn move to more efficient and greater performing equipment.
These improvements in equipment and systems come at a cost which may increase
the risk for fire. For example, higher operating temperatures or pressures can directly
change the range of conditions under which fire can occur for that system, and so too
the conditions under which fire-resistant fluids must perform in that system.
The proceedings of the Symposium on Fire Resistant Fluids were held at the
Fairmont Q.E. Montreal Hotel in Montreal, Quebec, Canada on June 24th, 2013. The
topics of the papers were related to fire-resistant hydraulic and compressor fluids and
liquid fuels. Specific chemistries discussed in one or more of the papers included
phosphate ester, polyol ester, polylalkylene glycol, water glycol, and thickened and
un-thickened high water fluids.
The editors wish to express their thanks and appreciation to the authors and
attendees of the symposium and to all the reviewers of these papers. We also owe a
debt of gratitude to the ASTM staff who were an essential part of both the successful
symposium proceedings and the publication of this volume.
John Sherman
Betsy Butke

vii
FIRE RESISTANT FLUIDS 1

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130179

John Sherman,1 Jonathon Maloy,3 Emidio Martino,3


Patrice Cusatis,2 and Paul Fasano2

Use of Thickened High Water


Hydraulic Fluid in Flat Rolled
Steel Production
Reference
Sherman, John, Maloy, Jonathon, Martino, Emidio, Cusatis, Patrice, and Fasano, Paul, “Use of
Thickened High Water Hydraulic Fluid in Flat Rolled Steel Production,” Fire Resistant Fluids,
STP 1573, John Sherman, Ed., pp. 1–23, doi:10.1520/STP157320130179, ASTM International,
West Conshohocken, PA 2014.4

ABSTRACT
Thickened HFA-E hydraulic fluid is in the class of most fire-resistant hydraulic
fluid as tested according to ISO 15029-2. A thickened HFA-E hydraulic fluid
demonstrated improved lubricity, load carrying capabilities, corrosion protection,
and bacterial resistance in comparison to the standard HFA-E (95/5) high water
hydraulic fluid while in operation in the roughing and finishing mill hydraulic
systems at the ArcelorMittal Dofasco steel production complex. The
improvements in hydraulic fluid properties and hydraulic system operation
resulted in increased system reliability, decreased maintenance costs, and
extended equipment life for the roughing and finishing mill hydraulic systems.

Keywords
fire-resistance, polyalkylene glycol, high water hydraulic fluids, HFA-E hydraulic
fluids, FM approvals

Manuscript received November 30, 2013; accepted for publication May 30, 2014; published online July 18,
2014.
1
BASF Corporation, Fuel and Lubricant Solutions, 1609 Biddle Avenue, Wyandotte, Michigan 48192.
2
BASF Corporation, Fuel and Lubricant Solutions, 500 White Plains Road, Tarrytown, New York, 10591.
3
ArcelorMittal Dofasco, 1330 Burlington Street East, Hamilton, ON, L8N 3J5, Canada.
4
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
2 STP 1573 On Fire Resistant Fluids

Introduction
CLASSIFICATION AND EVALUATION OF FIRE-RESISTANT
HYDRAULIC FLUIDS
Fire-resistant hydraulic fluids are used in industries where hydraulic systems are in
close proximity to high temperature surfaces or other sources of ignition, and
thereby a danger to human life [1–6]. These specialized fluids are also used where
mechanical fire suppression measures are not practical or in combination with
mechanical fire suppression systems for more comprehensive protection. In general,
for industrial applications where fire hazards exist, the use of fire-resistant hydraulic
fluids at some level is typically a requirement for obtaining insurance for the indus-
trial facility. Industries that require the use of fire-resistant hydraulic fluids include:
coal mining, steel manufacture, aluminum die-casting, automobile production,
commercial aviation, power generation, and military applications. All fire-resistant
fluids fall into two basic categories: those that contain water and those that do not
contain water [7]. The general designation for hydraulic fluids is HF [8]. Fire-
resistant hydraulic fluids are further classified by composition according to ISO
6734/4 [9] as shown in Table 1.
Today the most widely recognized standards for evaluation of fire-resistant
hydraulic fluids are ISO 12922: lubricants, industrial oils and related products (class
L)–Family H (hydraulic systems)—specifications for categories HFAE, HFAS, HFB,
HFC, HFDR, and HFDU [10], and the FM approval standard CN 6930 [11,12]. The
FM approval standard evaluates only the fire-resistance of the hydraulic fluid while
the ISO 12922 standard evaluates hydraulic fluid performance as well as fire-
resistance. FM Approvals Standard CN 6930 (April, 2009) [13] describes the flam-
mability classification of industrial fluids using two different methods: one method
for fluids having a fire point (non-water containing) by determining the spray flam-
mability parameter of the fluid and the other method for fluids not having a fire
point (water containing) by determining the adiabatic, stoichiometric flame temper-
ature of the fluid. Fluids that pass their respective tests are designated as “FM
approved.” FM approvals states that hydraulic fluids that contain greater than 80 %
water such as HFA-E and HFA-S types do not need to be tested to obtain the “FM
approved” designation. The expected results for the standard types of fire-resistant

TABLE 1 ISO 6743-4 classification of fire-resistant hydraulic fluids.

Fluid Classification Fluid Description

HFAE Oil-in-water emulsions, typically more than 80 % water content


HFAS Chemical solutions in water, typically more than 80 % water content
HFB Water-in-oil emulsions
HFC Water–polymer solutions, typically less than 80 % water
HFDR Water-free synthetic fluids consisting of phosphate esters
HFDU Water-free synthetic fluids of other compositions than HFD-R
SHERMAN ET AL., DOI 10.1520/STP157320130179 3

TABLE 2 Typical results for fire-resistant hydraulic fluid types evaluated according to FM approvals
standard CN 6930 (April 2009).

Fire-Resistant Hydraulic Fluid Types Typical Result of Evaluation

HFA FM approved
HFB FM approved (if water > 45 %)
HFC FM approved (if water > 30 %)
HFDR FM approved
HFDU FM approved
Mineral Oil Not approved

hydraulic fluids as outlined in ISO 6743-4 when tested according to the appropriate
protocol in the FM Approvals CN 6930 standard are summarized in Table 2.
ISO 12922: lubricants, industrial oils and related products (class L)—Family H
(hydraulic systems)—specifications for categories HFAE, HFAS, HFB, HFC, HFDR,
and HFDU provides the technical requirements for all classes of fire resistant
hydraulic fluids used in hydrostatic and hydrodynamic hydraulic systems for
general industrial applications. The standard was based on the technical and
fire-resistance requirements outlined in the seventh Luxembourg report [14]. The
flammability test methods required in ISO 12922 include ISO 15029-2, a stabilized
flame heat release method [15], which may be the most important fire resistant
spray ignition method within ISO 12922 because of its ability to reproducibly dis-
criminate levels of fire-resistance among types of fire-resistant hydraulic fluids
tested. The method determines the ignitability factor, RI, of the fluid. The ranges
for relative ignitability (RI) index grades in the ISO 15029-2 method are outlined in
Table 3. The typical RI index grades for common fire-resistant hydraulic fluid types
according to the ISO 15029-2 method, described by Phillips et al. [16], are shown in
Table 4. The lower the letter index grade the greater the fire-resistance of the fluid.
Another fire-resistance test required according to ISO 12922 is the Manifold

TABLE 3 Ranges for relative ignitability index grades according to ISO 15029-2.

Relative Ignitability Index Grades Range of Relative Ignitability Index

A >100
B 100–80
C 79–65
D 64–50
E 49–36
F 35–25
G 24–14
H <14
4 STP 1573 On Fire Resistant Fluids

TABLE 4 Typical relative ignitability index grades according to ISO 15029-2 for fire-resistant
hydraulic fluid types.

Fire-Resistant Hydraulic Fluid Types Typical Relative Ignitability Index Grade

HFA A
HFB E
HFC B–C
HFDR D–E
HFDU G–H
Mineral Oil H

ignition test ISO 20823 [17]. In this test method a 10 ml test sample of fluid is
dropped from a predetermined height and at a specific rate, onto a tube heated to
700 C, or another predetermined temperature (see Fig. 1). The resulting spray is
examined for flash or burn, both on the tube and after dripping from the tube on

FIG. 1 Equipment set up for Manifold ignition test ISO 20823.


SHERMAN ET AL., DOI 10.1520/STP157320130179 5

FIG. 2 Result of a mineral oil based hydraulic fluid tested according to ISO 20823 with
the tube heated to 700 C.

the pan below. An optimum result designation of “N” is reported when the
fluid does not flash or burn at any time. Figure 2 demonstrates the result of a
mineral oil based hydraulic fluid tested according to ISO 20823 with the tube heated
to 700 C.

HFA Type Fire-Resistant Hydraulic Fluids


HFA or high water content hydraulic fluids (HWCFs) have excellent fire resistance
properties combined with good to excellent corrosion protection but limited
lubrication properties [18–20]. Hydraulic equipment using this type of fluid must
be capable of operating with a low viscosity fluid. HFA type fluids are preferred
even in water hydraulic equipment as the HFA fluid will inhibit corrosion in
components and piping, control bacterial, and fungal growth and lubricate valves.
HFA type fluids are used in steel mills, tunnel boring equipment, over-ground pipe-
laying, and underground longwall mining equipment and other water hydraulic
equipment based systems [21,22].
6 STP 1573 On Fire Resistant Fluids

There are two sub classifications of the HFA category stipulated in ISO 6743-4.
One type is HFA-E: a macroemulsion or microemulsion typically blended on site
by blending 1 %–5 % of the HFA-E concentrate with water. Typically, HFA-E
microemulsions exhibit improved stability and anti-wear properties compared to
macroemulsions. The other HFA sub classification is HFA-S, an aqueous polymer
solution also blended from a concentrate on site. Both types have a minimum of
80 % water in the final operating fluids. The HFA-S type fluids are used with low
working pressures, where there is a possible ingression to sources of water and
where bacterial and fungal growth should be minimized. Applications with higher
working pressures require the significantly better lubricant properties of HFA-E
type hydraulic fluids. The HFA-E fluids typically have biocides added to the formu-
lation to control bacterial and fungal activity.
Reported vane and piston pump testing on HFA type fluids indicated a signifi-
cantly better wear resistance in an HFA-E microemulsion fluid in comparison with
an HFA-S solution fluid [23]. In fluid evaluations according to DIN 59389 Part 3,
only the HFA-E microemulsion fluid operated in the vane pump at 783 psi (54
bars) could complete the test duration of 250 h [23]. The HFA-E microemulsion
fluid vane pump tests operated at 1015 psi (70 bars) and 1378 psi (95 bars) were
aborted at approximately 180 h and 40 h, respectively, due to the development of
excessive noise and wear [23]. Testing of the fluid types on piston pumps for 1250 h
at a constant pressure of 2030 psi (140 bars) and alternating pressures of
1015/2030 psi (70/140 bars) indicated significantly less component wear (bearing
plate, reversing plate, and ball joint) with the HFA-E microemulsion fluid versus
the HFA-S solution fluid [23].
Thickened versions of The HFA-E fluids contain a percentage of synthetic
polymer thickener and are typically blended with 90 % water to form macro-
emulsions [24]. The polymeric thickeners used typically demonstrate an associative
thickening mechanism resulting from the interaction between hydrophobic
components of the polymer with other polymer and surfactant molecules. This
thickener/thickener and thickener/surfactant interaction results in associative com-
plexes with increased thickening ability.
A thickened HFA-E macro-emulsion fluid blended from a concentrate desig-
nated as BL-8022 was developed and evaluated at a concentration of 88.6 %
water. In Fig. 3, the fire-resistance of a HFA-E hydraulic fluid, which is com-
prised of 5 % fluid concentrate and 95 % water and is demonstrated during a
test according to ISO 20823 with the tube heated to 700 C. Figure 4 was taken
during the testing of the thickened HFA-E hydraulic fluid and, according to ISO
20823, comprised of 11.4 % BL-8022 concentrate and 88.6 % water with the tube
heated to 700 C. Both the HFA-E hydraulic fluid and the thickened HFA-E
hydraulic fluid shown in Figs. 3 and 4, respectively, achieved the optimum test
rating designation of “N.” The properties of the BL-8022 based HFA-E hydraulic
fluid are listed in Table 5. Typical HFA-E hydraulic fluids do not protect against
vapor phase corrosion inhibition; however, the thickened HFA-E hydraulic fluid
SHERMAN ET AL., DOI 10.1520/STP157320130179 7

FIG. 3 HFA-E hydraulic fluid; comprised of 5 % fluid concentrate and 95 % water, shown
during a test according to ISO 20823 with the tube heated to 700 C.

FIG. 4 Testing of the thickened HFA-E hydraulic fluid; comprised of 11.4 % BL-8022
concentrate and 88.6 % water, according to ISO 20823 with the tube heated to
700 C.
8 STP 1573 On Fire Resistant Fluids

TABLE 5 Typical properties for thickened HFA-E hydraulic fluid BL-8022.

Property Method Thickened HFA-E Hydraulic Fluid from BL-8022 Concentrate

Percent water ASTM D1744 88.6


Viscosity cSt at 40 C ASTM D445 46.9
Viscosity cSt at 10 C ASTM D445 351
pH at 25 C ASTM E70 10.3
Pour point  C ASTM D97 3
Air release in minutes ASTM D3427 30
Falex pin and vee block - ASTM D2670 2500
seizure load in lbs.
Copper corrosion ASTM D130 1B
Ferrous corrosion ASTM D665A Pass
Vapor phase corrosion BASF FF0002 Pass
Foam test sequence 1 ASTM D892 50/0

based on the BL-8022 concentrate does provide vapor phase corrosion


protection.
The BL-8022 based, thickened HFA-E fluid was vane pump tested according to
ASTM D2882-00 [25] but with revised temperature and pressure settings. Operat-
ing conditions according to ASTM D2882 [25] are for the fluid to be tested on a
Vickers 104C or 105C vane pump for 100 h at a temperature of 150 C, a speed of
1200 rpm, and a pressure of 2000 psi. The pressure was decreased to 1500 psi and
the operating temperature to 120 F for the vane pump test run on the BL-8022
based thickened HFA-E fluid. The result of the test on a Vickers 104C vane pump
was a total wear of ring and vanes combined of 9 mg. Evaluation of HFA-E and
HFA-S fire-resistant hydraulic fluid lubrication by vane and piston pump testing
was thoroughly described in Ref. [23]. The author of the article graphically reported
that the wear rate was approximately 20 mg per h for a 95/5 microemulsion hydrau-
lic fluid run in a vane pump at 1378 psi for approximately 40 h before the test was
aborted due to excessive noise.

HFA-E (95/5) Fluid Usage at ArcelorMittal


Dofasco
ArcelorMittal Dofasco is a supplier of high quality flat rolled steels, located in the
city of Hamilton at the western end of Lake Ontario on the St. Lawrence Seaway.
The company’s 750-acre steelmaking complex features state of the art facilities
that are among the most efficient, flexible and technologically advanced in North
America. These include three coke plants, two operating blast furnaces, a basic
oxygen steelmaking furnace, an electric arc furnace, two slab casters, a hot strip
SHERMAN ET AL., DOI 10.1520/STP157320130179 9

rolling mill, pickling lines, cold rolling mills, annealing and tempering facilities,
galvanizing lines, an electrolytic tinning line, and two tube mills. HFA-E (95/5)
hydraulic fluids had been used in the roughing mill and finishing mill hydraulic
systems since approximately 2000. The # 2 hot mill finishing and roughing mills
comprise the largest hydraulic system at the ArcelorMittal Dofasco complex. A
bulk reservoir containing 10 000 gal of hydraulic fluid feeds the finishing and
roughing mill reservoirs. The 7000 gal finishing mill reservoir supports the roll
balance, quick roll change, moving floors, crop shear, and coiler bank hydraulic
systems. The 2000 gal roughing mill reservoir supports the roll and spindle bal-
ance hydraulic systems.

Overview of Roughing Mill (RM) Finishing Mill


(FM) Hydraulic Systems
The finishing mill/roughing mill HFA-E hydraulic system used a blend of 95/5 city
water and HFA-E concentrate. The lubrication shop houses a 10 000 gal bulk
storage tank and proper mixing controls/transfer pumps to meet 90/10 require-
ments of both the FM and RM equipment. The main HFA-E bulk tank supplies
HFA-E fluid to the rougher and finisher reservoirs via a remotely controlled trans-
fer pump located in the lubrication shop. The fluid being transferred to the FM/RM
reservoirs must first pass through a flow meter, which monitors the amount of fluid
being added.
Located at the RM is a 1200 gal steel reservoir that supplies fluid for the
RM spindle and balance circuits. Make-up hydraulic fluid is provided manually
via a remote transfer pump, fill line, and a flow meter device that allows for
fluid management. The actual fluid level in the system reservoir is 1000 gal
during normal rolling and is monitored by level gauge. Three Triplex pumps
(one as a spare) supply the required flow to two on line weight loaded
accumulators.
The roughing mill (see Fig. 5) uses three hydraulic systems to operate the
following functions:
• Rougher chock clamps—1500 psig
• Rougher roll balance—1500 psig
• Rougher spindle balance—top and bottom, 760 psig and 1500 psig
• Rougher spindle thrust link balance—1500 psig
• Work roll change—2000 psig (mineral oil hydraulic fluid used here)

The Rougher functions are supplied by two of the three hydraulic systems.
The systems provide a hydraulic pressure of 760 and 1500 psig, respectively. The
top and bottom rougher spindle balance functions are designed to operate pri-
marily from the 760 psig system during rolling and from the 1500 psig system
while on roll change. All work roll change functions (roll extract, top work roll
off-setting W/R side shift and clevis hook lift) operate off the remaining 2000
psig system.
10 STP 1573 On Fire Resistant Fluids

FIG. 5 Single stand reversing rougher in roughing mill.

The low pressure hydraulic system services the spindle balance. It consists of 1
accumulator, 1 Triplex pump, and 1 standby pump that are shared with the
medium pressure system. Only one pump is in operation at a time. If an operating
pump fails, the standby pump can be selected to service the low pressure accumula-
tor by manually operating a valve to direct fluid flow to that accumulator. The
pump is started and stopped based on the accumulator level. If the pump was run-
ning and the stop button is pressed, it will be disabled until it has been reset by
manually pressing the start button for that pump. Five directional control valves are
used to control the circuits.
The medium pressure hydraulic system (1500 psig) services the roll balance. It
consists of 1 accumulator, 1 Triplex pump, and 1 standby pump that are shared
with the low pressure system. Only one pump is in operation at a time. If an operat-
ing pump fails, the standby pump can be selected to service the medium pressure
accumulator by manually operating a valve to direct fluid flow to that accumulator.
The pump is started and stopped based on the accumulator level. If the pump was
running and the stop button is pressed, it will be disabled until it has been reset by
manually pressing the start button for that pump. Three directional control valves
are used to control the circuits.
Located at the FM is a 3000 gal reservoir that supplies fluid for the FM QRC
equipment, seven stands having ten circuits per stand are each controlled by direc-
tional control valves (see Fig. 6). Crop shear and nineteen spray banks are controlled
by seven directional control valves. Make-up hydraulic fluid is provided manually
via a remote transfer pump, fill line, and a flow meter device that allows for fluid
SHERMAN ET AL., DOI 10.1520/STP157320130179 11

FIG. 6 Seven stands in finishing mill.

management. The actual fluid level in the system reservoir is 2550 gal during nor-
mal rolling and is monitored by one magnetic flag level gauge.
The reservoir serves several functions; the primary function is to contain all
the fluid used in the QRC and strip cooling hydraulic systems. A second func-
tion is to allow the fluid to be reconditioned on line. This is accomplished with
the use of a local kidney loop with filtration and cooling, a magnetic filter basket
and several air breathers (40 lm profile filters are used in the kidney loop filter
housing).
Centrifugal preload pumps supply hydraulic fluid from #1 tank at a charged
pressure of 40 psi to the suction side of the main pumps (#1-#4 Triplex pumps).
Pumps #5 and #6 (in line axial pumps) suction ports rely on #1 reservoir head pres-
sure to maintain a positive inlet pressure. Therefore, #5 and #6 pumps can be oper-
ated without the preload pumps. System working pressure is 1750 psig. Since the
accumulator weight is not precise nor is the mechanics of each accumulator exactly
similar, it would be impossible to control loading of both accumulators simultane-
ously. Therefore an accumulator blocking circuit is incorporated in to each ballast
accumulator, as logic control circuit.
Original design of the blocking circuit used hard limits and a timing circuit to
control switch over of the primary accumulator to secondary and back. The hard
limits have now been replaced with an ultrasonic level control (soft limits), which
provides greater flexibility in settings and also improves reliability. The ultrasonic
limit stages the QRC pumps to load or unload the pumps based on accumulator
height.
12 STP 1573 On Fire Resistant Fluids

TABLE 6 Reliability and maintenance for roughing and finishing mill hydraulic systems in 2008.

2008 Reliability and Maintenance for Roughing Mill


and Finishing Mill Hydraulic Systems Cost

Reliability—24 hours of delays $1 200 000 CD of lost production opportunity


$260 000 CD wasted energy cost of idle mill
Maintenance-valve and pump failures, high leakage
rates of bleed and directional valves resulting in $600 000 CD
longer charging times and wasted energy

Operation of RM and FM Hydraulic Systems


Using HFA-E (95/5) Hydraulic Fluid
The roughing and finishing mill hydraulic systems using the HFA-E hydraulic
fluid had a history of unreliability, high maintenance costs, and poor hydraulic
fluid quality. In 2007, there were 17 h production delays attributable to the rough-
ing and finishing mills operation resulting in $884 340 Canadian dollars (CD) of
lost production opportunity and $187 000 CD of wasted energy costs while the
mill was idle. Premature pump failures and corroded valves in addition to high
hydraulic fluid leakage rates resulted in extraordinarily high maintenance costs.
The reliability and maintenance costs for the two mill hydraulic systems in 2008
is summarized in Table 6. The condition of the HFA-E hydraulic fluid was identi-
fied as one of the main reasons for reliability and maintenance problems. The
HFA-E hydraulic fluid continued to have high bacteria levels putting plant per-
sonnel at risk despite the continued treatment of the systems with toxic bacteri-
cides in an effort to control the fluid bacteria levels. Fluid bacteria levels in the
systems were routinely above the 10 000 count alarm level despite weekly moni-
toring and treatment. The bacteria (Fig. 7), by ingesting hydraulic fluid additives,
compromised the fluid corrosion protection and in excreting acidic compounds
made it impossible for the pH of the fluids to be maintained above a pH of 8.5,
resulting in the systems always being in a potential corrosive state. Fluid cleanli-
ness was poor due to an inability to filter out the high levels of bacteria and
resulting gels causing high contaminant levels including metal particulates. The
low viscosity of the fluid—less than 1 cSt at 40 C—increased the internal leakage
and provided minimal lubrication to the pumps. Purchase of low viscosity valves
costing of 2 to 6 times as much as the current valves being used in the systems
were being considered as a last resort measure to lower the maintenance costs.
An investigation was initiated by ArcelorMittal Dofasco to identify alternative hy-
draulic fluids having the following properties:
• Low cost
• Good corrosion and wear properties
SHERMAN ET AL., DOI 10.1520/STP157320130179 13

FIG. 7 Bacterial growth on ceiling of finishing mill reservoir.

• Compatible with current HFA-E fluid


• Fire-resistance
• Viscosity greater than 20 cSt at 40 C
• Good filterability

In 2009, a hydraulic fluid was identified that potentially met their requirements:
the hydraulic fluid, a thickened HFA-E type produced from BL-8022 hydraulic fluid
concentrate.

Thickened HFA-E Fluid Introduction


at ArcelorMittal Dofasco
In July 2009, the initial volumes of thickened HFA-E BL-8022 concentrate were
transported to ArcelorMittal Dofasco to start the process of replacing the HFA-E
95/5 hydraulic fluid in the roughing mill and finishing mill hydraulic systems. A
comparison of the properties of the HFA-E (95/5) hydraulic fluid used in the
roughing mill and finishing mill hydraulic systems up to July 2009 and the thick-
ened HFA-E hydraulic fluid based on BL-8022 concentrate are summarized in
Table 7. The appearances of the two hydraulic fluids are shown in Fig. 8. It was esti-
mated it would take approximately a year to fully change out the HFA-E 95/5 fluid
14 STP 1573 On Fire Resistant Fluids

TABLE 7 Property comparison of thickened HFA-E hydraulic fluid based on BL-8022 concentrate
and original HFA-E (95/5) hydraulic fluids used in roughing and finishing mill hydraulic
systems at ArcelorMittal Dofasco.

Thickened HFA-E Hydraulic Original HFA-E (95/5)


Fluid from BL-8022 Hydraulic Fluids used at
Property Method Concentrate ArcelorMittal Dofasco

Percent Water ASTM D1744 88.6 95


Viscosity cSt at 40 C ASTM D445 46.9 0.8
Viscosity cSt at 10 C ASTM D445 351 166
pH at 25 C ASTM E70 10.3 10.0
Pour Point  C ASTM D97 3 6
Air Release minutes ASTM D3427 30 <30
Falex Pin and Vee Block ASTM D2670 2500 500
- seizure load in lbs.
Copper Corrosion ASTM D130 1B 2C
Ferrous Corrosion ASTM D665A Pass Pass
Vapor Phase Corrosion BASF FF0002 Pass Fail
Foam Test Sequence 1 ASTM D892 50/0 300/50

in the roughing mill and finishing mill systems in the course of normal operation to
the thickened HFA-E BL-8022 hydraulic fluid. It was determined the initial concen-
tration of the BL-8022 finished fluid would be 9 % of the BL-8022 concentrate in
91 % water during the replacement process. In July 2010, the concentration of the

FIG. 8 Rusted housing and swashplate from # 5 and # 6 pumps.


SHERMAN ET AL., DOI 10.1520/STP157320130179 15

FIG. 9 Bacterial count of samples taken from finishing mill reservoir at ArcelorMittal
Dofasco for the years from 2000 to 2013.

BL-8022 concentrate in the finished hydraulic fluid was increased to 11.4 %; the tar-
geted percentage to obtain the results required by ArcelorMittal Dofasco.

Bacterial Count, pH and Viscosity Results from


RM FM Hydraulic Systems Using Thickened
HFA-E Hydraulic Fluid
In the finishing mill, hydraulic system introduction of the thickened HFA-E
hydraulic fluid resulted in a significant decrease in bacterial activity. The non-water
portion of HFA-E hydraulic fluids is typically comprised predominantly of mineral
oil, which serves as a ready food source for bacteria. The BL-8022 thickened HFA-E
hydraulic fluid contains no mineral oil, but its non-water portion is composed of
predominantly polyalkylene glycol thickener, which is bio-resistant and does not
serve as a food source for bacteria. The bacterial counts recorded from samples
taken from the finishing mill reservoir from 2000 to 2013 are summarized in Fig. 9.
The roughing mill hydraulic system demonstrated a similar decrease in bacterial
activity after replacement with the thickened HFA-E hydraulic fluid. The bacterial
counts recorded from samples taken from the roughing mill reservoir from 2000 to
2013 are summarized in Fig. 10. When the roughing and finishing mill hydraulic
systems were using the HFA-E (95/5) hydraulic fluid, the high bacterial activity
16 STP 1573 On Fire Resistant Fluids

FIG. 10 Bacterial count of samples taken from roughing mill reservoir at ArcelorMittal
Dofasco for the years from 2000 to 2012.

made it impossible to maintain the fluid at a pH above 8.5 consistently. This is


where the potential for corrosion was minimal. Use of the thickened HFA-E
hydraulic fluid blended from BL-8022 concentrate in the finishing mill hydraulic
system resulted in the average fluid pH increasing from an average of less than 8.5
to an average of approximately 9.5 in samples taken from the finishing mill reser-
voir as illustrated in Fig. 11.
A similar improvement in pH was observed in the roughing mill hydraulic sys-
tem as demonstrated in Fig. 12. The initial thickened HFA-E hydraulic fluid replac-
ing the HFA-E (95/5) hydraulic fluid in the roughing and finishing mill systems
had a blend ratio of 9 % BL-8022 concentrate to 91 % water. This fluid increased
the fluid viscosity of the finishing mill hydraulic system to approximately 15 cSt at
40 C. In June 2010, the blend ratio of the thickened HFA-E hydraulic fluid used to
replenish the fluid in the system was revised to 11.4 % BL-8022 concentrate to
88.6 % water. This blend ratio, targeted to obtain optimum results, increased the
fluid viscosity of the finishing mill hydraulic system to approximately 40 cSt at
40 C. The viscosity increase over the time period of June 2009 to June 2013 for
samples taken from the finishing mill reservoir are demonstrated in Fig. 13. A vis-
cosity increase from approximately 1 cSt at 40 C to approximately 8 cSt at 40 C by
November 2010 was observed in the roughing mill hydraulic system. This system,
SHERMAN ET AL., DOI 10.1520/STP157320130179 17

FIG. 11 pH of samples taken from finishing mill reservoir at ArcelorMittal Dofasco for the
years from 2001 to 2012.

much smaller than the finishing mill system, was also replenishing less frequently;
therefore viscosity changes due to the use of the thickened HFA-E hydraulic fluid
took longer going by the calendar year. The sharp increase in viscosity due to
replenishment by the higher viscosity thickened HFA-E hydraulic fluid (11.4 %

FIG. 12 pH of samples taken from roughing mill reservoir at ArcelorMittal Dofasco for
the years from 2001 to 2012.
18 STP 1573 On Fire Resistant Fluids

FIG. 13 Viscosity of samples taken from finishing mill reservoir at ArcelorMittal Dofasco
for the years from 2009 to 2013.

BL-8022 concentrate to 88.6 % water) was not observed until 2011. The average
roughing mill system viscosity as observed from samples taken from the roughing
mill reservoir was approximately 22 cSt from 2011 to 2013. The reason the viscosity
in the roughing mill reservoir had not reached a similar viscosity to the finishing

FIG. 14 Viscosity of samples taken from roughing mill reservoir at ArcelorMittal Dofasco
for the years from 2009 to 2013.
SHERMAN ET AL., DOI 10.1520/STP157320130179 19

mill reservoir is not known at this time. The recorded viscosities over the time
period from June 2009 to June 2013 for samples taken from the roughing mill reser-
voir are shown in Fig. 14.

System Performance, Reliability and


Maintenance Cost Results from RM and FM
Using Thickened HFA-E Hydraulic Fluid
Immediately after the thickened HFA-E hydraulic fluid was charged to the roughing
mill and finishing mill systems, the leakage rates decreased due to the fluid viscosity
increase. Reduced wear rates in pumps and valves further reduced the leakage rates.
As of June 2013, the overall leakage rate has decreased by 80 % from the leakage
rates experienced prior to June 2009 when the systems were operating using the
HFA-E (95/5) hydraulic fluid.
Pumps #5 and #6 are in line tandem piston pumps that were installed as part of
the finishing mill system to add 2  106 dollars’ worth of capacity by reducing roll
change times. The life of the piston pumps which fail due to corrosion (see Figs. 8
and 15) was increased significantly with use of the thickened HFA-E hydraulic fluid
as summarized in Table 7.
The operational life of the 12 spray bank valves increased by 250 % resulting in
a savings of $30 000 CD. The valve upgrade project to purchase “raw “water valves
was now no longer necessary resulting in a savings of $100 000 CD. Maintenance
costs in 2012 were 55 % lower than that in 2008, a reduction of more than $300 000
CD as outlined in Fig. 16. Maintenance costs for 2013 are projected to be less than

FIG. 15 Rusted and worm shoes and pistons from # 5 and # 6 pumps.
20 STP 1573 On Fire Resistant Fluids

FIG. 16 Summary of system delays of finishing mill and roughing mill hydraulic systems
and the resultant lost opportunity costs from 2008 to 2013.

2012. System delays in 2012 were 45 % lower than the delays reported in 2008,
resulting in an additional $600 000 CD of production time potentially added in
2012. This information is summarized in Fig. 18. The systems’ delays are projected
to decrease further in 2013 as presented in Fig. 17.

FIG. 17 Summary of maintenance costs of finishing mill and roughing mill hydraulic
systems from 2008 to 2013.
SHERMAN ET AL., DOI 10.1520/STP157320130179 21

Conclusions
Thickened HFA-E hydraulic fluid has equivalent fire-resistance to HFA-E (95/5)
hydraulic fluids but has significantly improved performance in viscosity, lubricity,
corrosion protection, and bacterial resistivity.
At Arcelor Mittal Dofasco, the high water hydraulic fluid systems were stabi-
lized with improvements in the fluid properties to include:
• No bacterial growth
• Increased viscosity
• Improved fluid cleanliness due to improved fluid filterability
• Increased lubrication
• Maintained fire-resistance

The high water hydraulic systems have realized more than a million dollars in
savings since 2009 due to decreased maintenance costs and improved system reli-
ability as a result of using the thickened HFA-E hydraulic fluid. In summary, the
change from the HFA-E (95/5) hydraulic fluid to the thickened HFA-E hydraulic
fluid at ArcelorMittal Dofasco complex stabilized their systems using HFA-E hy-
draulic fluids that up to that point were not in control resulting in significant cost
savings and increased productivity and operator safety.

References

[1] Mathe, J., “Fire-Resistant Hydraulic Fluids for the Plastics Industry,” SPE J., Vol. 23, 1967,
pp. 17–20.

[2] Sullivan, J. M., “Wanted: Cheap, Safe, Enviro-Friendly Fluids,” Lubr. World, September,
1992, pp. 49–53.

[3] White, D. F., “The Unintentional Ignition of Hydraulic Fluids Inside High Pressure Pneu-
matic Systems,” ASNE J., Vol. 72, 1960, pp. 405–413.

[4] Polack, S. P., “Progress in Developing Fire-Resistant Hydraulic Fluids,” Iron Steel Eng.,
Vol. 35, 1958, pp. 87–92.

[5] Davis, R., “Fire-Resistant Hydraulic Fluids,” Q. J. NFPA, Vol. 52, 1959, pp. 44–49.

[6] Myers, M. B., “Fire-Resistant Hydraulic Fluids—Their Application in British Mines,” Col-
liery Guardian, Oct. 1977.

[7] Harrison, A. J., “Fire-Resistant Hydraulic Fluids—Their Development and Use in the Min-
ing Industry,” Potash Technology: Mining Processing, Maintenance, Transportation, Occu-
pational Health and Safety, Environment, Pergamon Press, Oxford, UK, 1983, pp.
459–465.

[8] Commission of the European Communities, Safety and Health Commission for the Min-
ing and Extractive Industries, Working Party Rescue Arrangements, Fires and Under-
Ground Combustion, Committee of Experts on Fire-Resistant Fluids, “Sixth Report on
22 STP 1573 On Fire Resistant Fluids

Specifications and Testing Conditions Relation to Fire-Resistant Hydraulic Fluids Used


for Power Transmission (Hydrostatic and Hydrokinetic) in Mines,” Doc. 2786/8/81 E, Lux-
embourg, 1983.

[9] ISO 6734/4: Lubricants, Industrial Oils, and Related Products, Class LJ—Classification—
Part 4: Family H–Hydraulic Systems, ISO, Geneva, Switzerland.

[10] 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,
Geneva, Switzerland, 2013.

[11] Factory Mutual Research Corporation, Approval Standard: Less Hazardous Hydraulic
Fluid, Factory Mutual Research Corp., Nonvood, MA, 1975.

[12] FM Approvals: Approval Standard: Flammability Classification of Industrial Fluids (Class


6930), Factory Mutual Global, Johnston, RI, 2002.

[13] FM Approvals: Approval Standard: Flammability Classification of Industrial Fluids (Class


6930), Factory Mutual Global, Johnston, RI, 2009.

[14] Luxembourg Commission of the European Economic Communities, “Requirements and


Tests Applicable to Fire Resistant Hydraulic Fluids used for Power Transmission and
Control,” L-2920m, DG\E\4, The 7th Luxembourg Report, Doc No. 4746/10/91,
Luxembourg, 1994.

[15] ISO 15029-2: Petroleum and Related Products—Determination of Spray Ignition Charac-
teristics of Fire-Resistant Fluids- Part 2—Spray Test-Stabilized Flame Heat Release
Method, ISO, Geneva, Switzerland, 2012.

[16] Phillips, W. D., Goode, M. J., and Winkeljohn, R., “Fire-Resistant Hydraulic Fluids and the
Potential Impact of New Standards for General Industrial Applications,” 100-1.12,
National Fluid Power Association, Milwaukee, WI, 2000.

[17] ISO 20823: Petroleum and Related Products—Determinations of the Flammability Char-
acteristics of Fluids in Contact With Hot Surfaces – Manifold Ignition Test, ISO, Geneva,
Switzerland, 2003.

[18] Coker, C. T. and Francis, C. E., “The Place for Emulsions as Fire-Resistant Power Trans-
mission Fluids,” Lubr. Eng., Vol. 12, No. 5, 1956, pp. 323–326.

[19] Deakin, P., “Fire Resistant Hydraulic Fluids,” Mining Technol., November/December,
1990, pp. 300–303.

[20] Garti, N., Felkenkrietz, R., Aserin, A., Ezrahi, E., and Shapira, D., “Hydraulic Fluids
Based on Water-in-Oil Microemulsions,” Lubr. Eng., Vol. 49, No. 5, 1993, pp.
404–411.

[21] Brooke, B., “Development of a High-Water-Based Fluid System for a Universal


Beam Rolling Mill,” Conference Proceedings: Hydraulics, Electrics and Electronics in
Steel Works and Rolling Mills, Mannesmann Rexroth GmbH, Lohr-am-Main, Germany,
1994.

[22] Young, K. J. and Kennedy, A., “Development of an Advanced Oil-in Water Emulsion Hy-
draulic Fluid, and it’s Application as an Alternative Mineral Oil Hydraulic Oil in a High
Risk Environment,” Lubr. Eng., Vol. 49, No. 11, 1993, pp. 873–879.
SHERMAN ET AL., DOI 10.1520/STP157320130179 23

[23] Janko, K., “A Practical Investigation of Wear in Piston Pumps Operated with HFA Fluids
with Different Additives,” J. Synth. Lubr., Vol. 4, 1987, pp. 99–114.

[24] Rasp, R. C., “Water-Based Hydraulic Fluids Containing Synthetic Components,” J. Synth.
Lubr., Vol. 6, 1989, pp. 233–252.

[25] ASTM D2882-00: Standard Test Method for Indicating the Wear Characteristics of
Petroleum and Non-Petroleum Hydraulic Fluids in Constant Volume Vane Pump
(Withdrawn 2003), Annual Book of ASTM Standards, West Conshohocken,
PA, 2000.
FIRE RESISTANT FLUIDS 24

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130177

Steven R. Westbrook,1 Bernard R. Wright,1


Steven D. Marty,1 and Joel Schmitigal2

Fire Resistant Fuel for Military


Compression Ignition Engines
Reference
Westbrook, Steven R., Wright, Bernard R., Marty, Steven D., and Schmitigal, Joel, “Fire Resistant
Fuel for Military Compression Ignition Engines,” Fire Resistant Fluids, STP 1573, John Sherman,
Ed., pp. 24–37, doi:10.1520/STP157320130177, ASTM International, West Conshohocken, PA 2014.3

ABSTRACT
During an Army research program in the mid-1980s, fire-resistant diesel fuel that
self extinguished when ignited by an explosive projectile was developed.
Chemically, this fire resistant fuel (FRF) was a stable mixture of diesel fuel, 10 %
water, and an emulsifier. The Army FRF program ended in 1987 without fielding
this fire resistant fuel formulation. There were both technical and logistical
reasons for this. Unconventional warfare experienced in Iraq and Afghanistan
involving use of Improvised Explosive Devices (IED) has led the Army to restart
the FRF program in an attempt to counter the increasing threat of fuel fires.
Efforts are now underway to develop new FRF to reduce and/or eliminate both
the initial mist fireball and any residual pool burning. Vehicle operation and
environmental conditions commonly cause the temperature of the fuel in the
vehicles to rise above its flash point, thus making it more susceptible to being
ignited. This elevated fuel temperature, when combined with an ignition source such
as a ballistic penetration near the fuel tank or fuel line, significantly increases the
potential for a catastrophic fuel fire. This paper will discuss some of the aspects and
limitations of developing a fire resistant fuel water emulsion and how the use of JP-
8, as intended by the single fuel forward concept, affects this development.

Keywords
diesel fuel, fire resistant fuel, water-emulsified fuel

Manuscript received November 27, 2013; accepted for publication May 8, 2014; published online July 22,
2014.
1
TARDEC Fuels and Lubricants Research Facility at SwRI, San Antonio, TX, 78238.
2
US Army RDECOM TARDEC, Warren, MI 48397.
3
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
WESTBROOK ET AL., DOI 10.1520/STP157320130177 25

Background
During an Army research program in the mid-1980s, fire-resistant diesel fuel that
self extinguished when ignited by an explosive projectile was developed. This fire
resistant fuel (FRF) was a stable mixture of diesel fuel, 10 % purified water contain-
ing less than 50 ppm dissolved solids, 6 % emulsifier, and 6 % aromatic hydrocar-
bon concentrate to aid in the solubility of the emulsifier [1].
Previous research, including the program headed by the Army in the 1980s,
involved using a variety of approaches to reduce the flammability of fuel. These
approaches evaluated emulsified fuel, halogenated additives, mist control addi-
tives, and water-in-fuel emulsions, the latter showing the most promise, for
ground vehicle applications [1,2]. In a water-in-fuel emulsion, water molecules
are suspended in fuel by the hydrophilic end of a surfactant which has its hydro-
phobic end dissolved in the base fuel. This fire resistant fuel was a clear to hazy
emulsion consisting of water, emulsifier premix (equal amounts of the emulsifier
and an aromatic concentrate), and diesel fuel. This emulsion performed satisfacto-
rily both in diesel and turbine engine systems and could be prepared in the field
for availability as needed. Although this earlier version of FRF did not eliminate
the initial mist fireball that occurs when a projectile impacts the vehicle, it signifi-
cantly reduced the fuel fire threat by retarding the flame-spread rate and would
self extinguish spilled fuel eliminating residual pool burning. The self-
extinguishing characteristic resulted from the heat sink provided by the water,
emulsified water on the surface of the fuel preventing fuel vaporization, and the
released water vapor that concentrating at the surface of the fuel eliminating oxy-
gen from the fuel.
By 1987, the urgency for the development of a fire resistant fuel had diminished
which resulted in the reallocation of funding. Additionally, there were both techni-
cal and logistical reasons for this. Filter plugging caused by the fuel at low tempera-
tures and the need for high purity water (beyond the purity obtained by typical
Army water purification systems) to ensure a stable emulsion were a few of the
major technical hurdles that the FRF program was unable to clear. The logistical
burden requiring the 12 % additive solution to create the FRF was also an obstacle.
Because of a combination of these problems associated with FRF, further efforts to
pursue this fuel were discontinued.

FRF Development Project


With the start of the conflicts in Iraq and Afghanistan, attention once again
returned to the fuel fire threat that was taking its toll on both vehicles and person-
nel. The Army uses JP-8 aviation fuel in ground vehicle operations during combat
situations as intended by the single fuel in the battlefield policy, as directed by the
DoD Directive 4140.43 entitled “Fuel Standardization,” which mandates the use of
JP-8 for air and ground forces. The shift to JP-8 enabled the Air Force and the
26 STP 1573 On Fire Resistant Fluids

Army to standardize on one single fuel for all operations. While the Air Force made
this move, among other reasons, to increase safety by moving away from JP-4, the
Army’s move to utilize JP-8 was a move toward a more volatile fuel, with a lower
flashpoint than Diesel 2 fuel that was used previously.
JP-8 is a kerosene-based fuel containing a distribution of hydrocarbons with
between 8 and 16 carbon numbers and having a minimum flashpoint of 38 C
(100 F) [3]. Diesel by comparison is a distillate fuel composed of a mixture of
hydrocarbons with between 12 and 21 carbon atoms per molecule giving it a mini-
mum flashpoint temperature of 52 C (125 F) [4]. The flashpoint and light end
components difference between diesel and JP-8 has become a large obstacle to over-
come in the development of a fire resistant JP-8 formulation that will self extinguish
when the fuel temperature is elevated to desert conditions, 65 C (149 F). The
higher volatility of the JP-8 fuel when ignited at desert conditions allows for a
higher proportion of the fuel to be ignited, and this prevents the water emulsion
from extinguishing the fire.
The U.S. Army uses compression ignition (diesel) engines to power a large
majority of the ground vehicles in both its tactical wheeled and combat vehicles
fleets. Most of these engines utilize fuel as a cooling agent for the engines fuel injec-
tor system and have a fuel delivery system that returns a portion of the fuel
from the injectors back to the fuel tank. This recirculation heats the fuel, commonly
raising the temperature of the fuel in the tank above its flash point, the lowest tem-
perature that the vapor above the fuel will ignite when exposed to an ignition
source, making the fuel more susceptible to being ignited.
The heating of the fuel used in compression ignition engines, when combined
with any direct or indirect ballistic penetration near the fuel tank or fuel line, signif-
icantly increases the potential for a catastrophic fuel fire. Having a fuel that would
not ignite under these conditions would have obvious benefits in terms of both
increased personnel and vehicle survivability.
In April 2007, a more comprehensive effort was initiated that involved the
following tasks:
• developing new emulsified fuel formulations,
• investigating mist control additives to diminish the fuel mist fireball,
• determining the effect of FRF on vehicle and equipment systems,
• designing a blending system for producing the FRF in the field,
• determining overall effectiveness of the FRF based on JP-8.

Initially, a new baseline had to be established (for blending and flammability)


using JP-8, as the previous work only evaluated diesel fuel. The development of
an emulsified fuel formulation that yields a stable emulsion (i.e., one that does not
separate) using JP-8 (or diesel fuel) has been the most difficult of all the
above tasks. Variables such as fuel composition, aromatic content, water quality,
emulsifier/surfactant chemistry, additive interactions, etc., need to be understood
and optimized. Adding to the complexity of this task is the addition of mist control
additives (long chained, high molecular weight polymers) to the emulsified fuel
WESTBROOK ET AL., DOI 10.1520/STP157320130177 27

formulation. The long chain polymers act to control fuel mist droplet size and thus
reduce the size of the initial fireball that occurs [5–7].
This paper is meant to give an overview of the main development areas
associated with formulating an optimal FRF. The areas include fuel fire resistance,
equipment performance impacts, and also fuel stability and applications. Differen-
ces between a JP-8 and a Diesel 2 fuel will also be discussed. A more detailed report
of the results is also available [8].

Results
FIRE RESISTANCE
The vehicle fuel fires experienced in combat situations occur in two distinct phases.
The first phase is commonly termed a fireball and seen as a fuel explosion. The
fireball phase is caused by the explosive or ordnance rupturing the fuel tank
and performing a rapid mechanical mixture of fuel spray and air mixture, which
combined with the head from the explosive or ordnance manifests itself as an
explosion. The second phase is the ignition and flame spread over the pool of fuel
spilt on from the vehicles fuel tank. The pool fire is caused by the pool of fuel
having a sufficient enough temperature to emit enough fuel vapors into the air
above the pool surface to sustain a fire.
Due to the increased volatility of the JP-8 fuel when compared to diesel fuel, it
is imperative to suppress both phases of these fires as water alone has shown it is
not capable of providing sufficient extinguishment at acceptable concentrations of
water, as was seen in the precursor work centered around diesel fuel. Therefore the
goal of the development of a fire resistant JP-8 is to minimize both phases of these
fuel fires.
Ballistic testing of different mist control additives incorporated into the emulsi-
fied fuel formulations was conducted. Figure 1 shows a side-by-side photo sequence
of a regular JP8 and a diesel based FRF fuel. In order to simulate battlefield fuel
tank conditions in a worst case/hot environment, the ballistic tests are being con-
ducted with the FRF pre-heated to 150 deg F. A 55 gallon steel barrel filled with
30 gallons of FRF test fuel is used as the test article. The pre-heating and 25 gallons
of vapor space simulate worst-case threat evaluation.
In the photo sequence in the left column, an untreated fuel displays typical fuel
fire behavior exhibiting the initial fireball caused by the fuel mist explosion, the pro-
ceeds though the flame propagation stage of the pool fire to conflagration. Within a
similar time frame, an FRF (containing water and mist control additive) in the right
column demonstrates a fire ball that is reduced in size by over 20 %, and self sup-
presses preventing the flame propagation as seen in the untreated fuel.
While the pictures in Fig. 1 show testing in a blast bunker, present testing is
conducted in the open, where there are no “bunker effects” with respect to oxygen
starvation or wind.
28 STP 1573 On Fire Resistant Fluids

FIG. 1 Conventional JP8 compared to diesel FRF.

In order to quantify the “fire resistance” effectiveness of a particular FRF


formulation during ballistics testing, a data acquisition system is used to record
temperature versus time measurements. The system consists of 10 thermocouples
(2 feet apart) spread out linearly across the testing area (just above the 30 gallon
steel barrel). Thermocouple 1 is directly above the fuel container and the remaining
thermocouples are increasingly farther away toward the front of the facility. Tem-
perature response during testing is recorded at a logging rate of 5 kHz for a total of
30 s. This information allows for the determination of the flame propagation rate
and severity of the initial fireball and resulting burn where applicable. Figure 2 below
WESTBROOK ET AL., DOI 10.1520/STP157320130177 29

FIG. 2 JP-8 ballistic test—uncontrolled burn.

shows a representative plot of an uncontrolled burn. Using this temperature and


time data, work was conducted to evaluate optimal FRF formulations exposed to
various threat types. Examination of Fig. 2 shows that the initial fireball was directly
beneath thermocouple 1 (at the fuel drum) but tended to shift as the fire progressed.
The shifting pattern of highest temperature can be attributed to the effects of wind
and fuel pooling, among others. The most important conclusion from these data is
the documented size and temperature of the fire/pool.
In contrast to Figs. 2, 3 is a representation of FRF, in this case a base fuel of Die-
sel No. 2 was utilized. Results from FRF made with diesel fuel were shown here in
order to demonstrate the dramatic difference that can be possible. The plots in Fig.
3 show that diesel-FRF can produce a significantly smaller initial fireball and will
rapidly self-extinguish. Similar tests with FRF made with JP-8 tended to produce a
less-dramatic difference in fireball and self-extinguishment.
To combat the fireball phase, fuel formulation work tests were conducted to
evaluate the mitigation properties of several mist control additives. The long chain
polymers act to control fluid droplet size by imparting non-Newtonian properties
into the fuel which decrease the surface to volume ratio and thus reduce the size of
the initial fireball that occurs [5–7]. Testing showed that these additives were able
to reduce fireball effects by reducing temperatures over 100 C, reducing fireball
duration from over 2.5 to 1.5 s, and an average 16 % reduction in fireball size. These
long chain, high molecular weight additives can reduce the initial fireball. The large
30 STP 1573 On Fire Resistant Fluids

FIG. 3 Diesel 2 FRF ballistic test—controlled burn.

molecules will shear down when exposed to the high pressure injection systems of
modern diesel engines that re-circulates a portion of the fuel back to the vehicles
fuel tank, reducing effectiveness as the vehicle completes its mission. Engine testing
demonstrated that mist control additive will degrade with successive passes through
the engine fuel system. While the degraded polymer is still 1–3 orders of magnitude
higher in average molecular weight than fuel molecules, its efficacy as a mist control
additive is certainly reduced. Research is underway to develop polymer mist control
additives that will be resistant to shear.
After the initial fireball, the water emulsion works to extinguish the fuel pool
fire by means of the heat sink, prevention of fuel vaporization, and elimination of
oxygen as described earlier. The mist control additives do not provide any fire
suppression properties in the fire pool phase. Because JP-8 is already at a power dis-
advantage, in terms of vehicle performance, when compared to Diesel 2, formulat-
ing in only “just enough” water is a critical design criteria.
The JP-8 base fuel used in the formation of the FRF has a significant impact
on the ability of the fuel to self extinguish. The JP-8 fuel specification (MIL-DTL-
83133F) calls for a minimum flashpoint from 38 C, but can range up to the high
60 s. Depending upon the refining process used to make the fuel, the flash point
will vary within this range. Because of the wide range of acceptable JP-8 flash-
point temperatures, the FRF formulation must be designed to perform on the
lowest flashpoint fuels encountered. Utilization of a base fuel with as high of a
flashpoint as possible allows for greater extinguishment characteristics to be
imparted.
WESTBROOK ET AL., DOI 10.1520/STP157320130177 31

EQUIPMENT PERFORMANCE IMPACTS


Engine dynamometer testing was done using multiple engine families commonly
used in Army vehicles. Shown below is data derived from the GEP 6.5 L(T) engine.
This is the engine used in the HMMWV (Hummer) and is the highest-density
engine in the Army fleet. Testing was conducted to look at engine horsepower, tor-
que and fuel consumption. The following three charts (Figs. 4–6) show the engine
performance effects experienced by utilizing JP-8 FRF based fuels. In summary, it
can be seen that the addition of water to the JP-8 fuel lowers the maximum torque
and horsepower while increasing fuel consumption. This is not unexpected as any
addition of water to the fuel lowers the overall energy content of the fuel. The mist
control additive (AMA) does add a very small amount of energy back into the fuel,
but it is nearly negligible. See Figs. 4–6.
From the above charts, an FRF based fuel with 10 % water, 250 ppm mist con-
trol additive would be expected to provide roughly 8 %–9 % less power, torque, and
fuel economy than neat JP-8.
Engine data was used in existing vehicle models to provide a prediction of the
effects of the FRF impacts on overall vehicle performance. The models showed the
vehicle response to the new fuel to be minimal, it would be expected that individual
vehicle operators may not notice the difference in fuel energies. Table 1 details the
loss in acceleration and vehicle to speed on a flat surface, while Table 2 details the
loss in speed on a slope.

FRF FUEL STABILITY AND PER USE MIXING


After a comprehensive initial evaluation of available fuel emulsifiers (including
polyalkoxylated phenols, polyalkoxylated alkyl phenols, polyethoxylated esters,
sorbitan esters, quarternary ammonium salts of fatty acids, cocoamidobetaine, and

FIG. 4 6.5 L turbo diesel maximum power output.


32 STP 1573 On Fire Resistant Fluids

FIG. 5 6.5 L turbo diesel maximum torque output.

FIG. 6 6.5 L turbo diesel brake specific fuel consumption.

TABLE 1 Simulated data for acceleration and vehicle to speed on a flat surface.

JP8 JP8 FRF

0–30 mph (s) 11.5 12.52


0–50 mph (s) 32.24 37.87
Top Speed (mph) 67 61
WESTBROOK ET AL., DOI 10.1520/STP157320130177 33

TABLE 2 Simulated data for vehicle to speed on a slope.

JP8 JP8 FRF

5 % grade (mph) 40 37
20 % grade (mph) 11 10
40 % grade (mph) 8 7
60 % grade (mph) 5 4

ethanolamides of fatty acids), testing was conducted on a refined list of emulsifiers


to find those that will produce a stable emulsion with water using a range of both
diesel and JP-8 fuels. Special care was taken to insure that the candidate emulsifiers
were not as sensitive to water quality as in the prior project work.
Emulsions can be broadly segregated in two groups: micro and macro emul-
sions. These groups differ by the size of the suspended water droplets. Most of the
emulsions presently being evaluated are in a class known as “micro-emulsions” due
to the increased stability experienced with this type of emulsion. These mixtures are
clear and bright. See Fig. 7.
Some FRF emulsions do not always remain clear/transparent, but rather appear
white and milk-like. See Fig. 8. The functional performance of a non-stratified
opaque FRF mixture is equivalent to clear FRF.
For this effort, emulsion stability was defined by the absence of any distinct
layers in the FRF mixture. Testing was conducted to statistically optimize and quan-
tify FRF emulsion stability. Variables included: temperature (hot or cold), base fuel,

FIG. 7 FRF mixtures—clear.


34 STP 1573 On Fire Resistant Fluids

FIG. 8 FRF mixtures—cloudy.

amount and type of emulsifier, amount and quality of water, and amount and type
of mist control additive. This testing included extended hot and cool storage, mate-
rial compatibility studies and eventually temperature cycling. It should be noted
that some stratified samples re-mixed with minor agitation, but most did not. An
oleamide diethanolamine additive was ultimately selected for continued work.
Emulsion stability for indefinite periods of time is desired, but not likely. There-
fore, with limited stability and operational use limits, the present expected deploy-
ment is FRF blending at re-fueling points where vehicles are fuel prior to high
threat missions. The requirements for this preliminary design were to maximize use
of existing Army petroleum and water handling equipment already available within
the inventory. Different configurations of the necessary pumping and mixing equip-
ment were considered, as detailed in Fig. 9. However, use of a dedicated pump per
fluid (i.e., water, fuel, and emulsifier/additive) that forces each through a static
mixer was also evaluated and eventually adopted.
At the onset of this project, one of the goals was to develop a formulation that
would produce a stable FRF with water that had up to 1000 ppm of dissolved solids.
Initial attempts to meet this goal centered on finding an emulsifier that would pro-
duce the desired results. Unfortunately, none of the emulsifiers that we evaluated
were found to produce an acceptable emulsion with water containing 1000 ppm sol-
ids. After an emulsifier was selected, we looked for alternative ways to reach the
water hardness goal. Variations of mixing time and energy were investigated with-
out success. Looking for yet another approach, we decided to mix a chelating agent
(ethylenediamine tetraacetic acid, EDTA) with the blend. The EDTA chelates the
solids and prevents them from inferring with the emulsifier. Through a series of
experiments, we found that adding EDTA at a 1:1 ratio, on a ppm basis with the
WESTBROOK ET AL., DOI 10.1520/STP157320130177 35

FIG. 9 Preliminary FRF field blending concept diagram.

measured solids in the water, yielded a stable emulsion. We obtained acceptable


results up to 1250 ppm of dissolved solids. We did not attempt mixing with water
above 1250 ppm solids.

Discussion
The Army’s utilization of JP-8 aviation fuel for ground fuel applications has a tre-
mendous impact on the development of a fire resistant fuel. The low flashpoint of
the JP-8 results in requiring the FRF to control both the fireball and pool fire phases
of vehicle fuel fires to be considered successful. Utilization of higher flashpoint base
fuels increases the ability of the FRF to be able to self-extinguish.
While desirable for FRF to have no performance impacts on vehicle operation,
the data produced within this research program has provided similar results as pre-
vious FRF research programs. While power and range losses are undesirable, they
are unavoidable and expected as any addition of water to the fuel lowers the overall
energy content of the fuel. While the drawbacks of these effects may preclude the
use of FRF in all ground vehicles at all times, there may be appropriate times when
commanders would see benefits in the use of FRF. Utilization of a base fuel with
greater energy density will offset the power loss experience by FRF JP-8.
Indefinite FRF stability is greatly desired, but the utilization of water in fuel
emulsions makes it impracticable at this time. Use of FRF at temperatures below
36 STP 1573 On Fire Resistant Fluids

the freezing point of water can be problematic as the water may separate or freeze.
Efforts are being made to increase the stability of FRF fuels and make them usable
across a greater range of environmental applications. However, it is not realistic to
expect that an FRF can be developed that will be useable in all regions of the world.

Summary, Conclusions and Recommendations


The difference in flashpoint and light end components between diesel and JP-8
has become a large obstacle to overcome in the development of a fire resistant JP-8
formulation that will self extinguish when the fuel temperature is elevated to desert
conditions, 65 C (149 F). JP-8 is a kerosene-based fuel having a minimum flash-
point of 38 C (100 F) while diesel No. 2 fuel by comparison is a distillate fuel
with a minimum flashpoint temperature of 52 C (125 F). The higher volatility of
the JP-8 fuel when ignited at desert conditions, allows for a higher proportion of
the fuel to be ignited which prevents the water emulsion, traditionally used in fire
resistant fuel, from extinguishing the fire.
FRF blends made with higher flash point fuels (diesel fuel) performed consis-
tently better in emulsion stability and flammability testing compared to blends
made with lower flashpoint fuels such as JP-8. Diesel fuel based FRF is available
today, with above identified issues, but JP-8 based FRF needs further work.
Low-temperature stability of emulsions continues to be a concern. Some emul-
sions, depending on fuel and water quality, maintained stability to several degrees
below 0 C. But most emulsions tended to break at about this temperature. We were
able to recombine the components with minimal mixing but the emulsion did not
have the same, typical, clear appearance as emulsions prior to freezing.
Under the right environmental conditions, FRF could be used in all of the
Army’s ground vehicles and support equipment, or just vehicles that are used in
high risk operations where exposure to IEDs is high.

References

[1] Wright, B. R. and Kanakia, M. D., “Final Review of U.S. Army Fire Resistant Fuel
Program,” Interm Report BFLRF No. 244, U.S. Army Belvoir Research, Development and
Engineering Center, Materials, Fuels, and Lubricants Laboratory, Fort Belvoir, VA, 1987.

[2] Wright, B. R., “Assessment of Concepts and Research for Commercial-Aviation Fire-Safe
Fuel,” Final Report SwRI No. 02-1800, Southwest Research Institute, San Antonio, TX, 2000.

[3] Department of Defense, “Turbine Fuels, Aviation, Kerosene types, NATO F-34 (JP-8),
NATO F-35, and JP-8þ100 (NATO F-37),” Military Specification MIL-DTL-83133F, Depart-
ment of Defense, Washington, D.C., 2008.

[4] ASTM D975-05: Standard Specification for Diesel Fuel Oils, Annual Book of ASTM
Standards, ASTM International, West Conshohocken, PA, 2005, www.astm.org, DOI:
10.1520/D0975-05.
WESTBROOK ET AL., DOI 10.1520/STP157320130177 37

[5] Mannheimer, R. J., “Rheological and Mist Ignition Properties of Dilute Polymer
Solutions,” Chem. Eng. Commun., Vol. 19, 1983, pp. 221–241.

[6] Mannheimer, R. J., “Real-Time Quality Control of Antimisting Kerosene (AMK),” Final
Report DOT/FAA.CT-85/5, FAA, Washington, D.C., 1985.

[7] Wright, B. R. and Weatherford, W. D., “Investigation of Fire-Vulnerability-Reduction


Effectiveness of Fire-Resistant Diesel Fuel in Armored Vehicular Fuel Tanks,” Final
Report AFLRL No. 130, U.S. Army Belvoir Research, Development and Engineering
Center, Materials, Fuels, and Lubricants Laboratory, Fort Belvoir, VA, 1980.

[8] Westbrook, S. R., Frame, E. A., Wright, B. R., Marty, S. D., Brandt, A. C., Hansen, G.,
Warden, R. W., Hutzler, S. A., and Johnson, J. E., “Fire Resistant Fuel,” Interim Report
TFLRF No. 416, U.S. Army TARDEC, Warren, MI, 2014.
FIRE RESISTANT FLUIDS 38

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130102

W. D. Phillips,1 J. W. G. Staniewski,2 and S. Suryanarayan3

Ion Exchange and Mechanical


Purification of Fire-Resistant
Phosphate Ester Fluids Used in
Steam-Turbine Control Systems
Reference
Phillips, W. D., Staniewski, J. W. G., and Suryanarayan, S., “Ion Exchange and Mechanical
Purification of Fire-Resistant Phosphate Ester Fluids Used in Steam-Turbine Control Systems,”
Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 38–74, doi:10.1520/STP157320130102,
ASTM International, West Conshohocken, PA 2014.4

ABSTRACT
Steam turbines at nuclear stations have electro-hydraulic control (EHC) systems
that use a phosphate ester-based fire-resistant fluid. This fluid undergoes
degradation in service via hydrolytic, oxidative, and thermal mechanisms that are
influenced by system design and operating conditions. Past experience (OPEX) has
shown that the condition of the fire-resistant fluid in service is critical for station
safety and nuclear regulatory authorities; therefore, chemistry control of this fluid is
included as a part of a station’s operating license. The typical industry approach to
maintaining fluid quality within specification is to continuously circulate a portion of
the fluid through an adsorbent solid to remove degradation products. Since the late
1980s, ion-exchange treatment has become one of the most effective purification
processes. However, there are now several different resin types available that can
interact with the fluid in different ways, and the optimum process for resin
treatment of phosphate esters has still to be identified. In fact, it will probably be
necessary to have several different options depending on the operating conditions.
EHC fluid purification is not limited to acidity control. It is also important to keep the
fluid clean and dry if it is to operate efficiently and offer a long service life.

Manuscript received June 9, 2013; accepted for publication October 3, 2013; published online July 16, 2014.
1
W. David Phillips and Associates, Stockport, Cheshire, United Kingdom.
2
Ontario Power Generation, Nuclear Services, 889 Brock Rd., Pickering, Ontario, Canada, L1W
3J2 (Corresponding author), e-mail: george.staniewski@opg.com
3
Kinectrics Inc., Chemistry, Toronto, Ontario, Canada.
4
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
PHILLIPS ET AL., DOI:10.1520/STP157320130102 39

Mechanical techniques are, therefore, needed to complement and maintain the


activity of the resin treatment. For example, resin fouling by particulates can reduce
its activity and this may require improved filtration. The main objective of this paper
is to present the initial results of a new comparison of resin behavior intended to
improve performance of the ion-exchange treatment at CANDU (Canada Deuterium
Uranium) nuclear stations. Also included are the results of early investigations into
different techniques for drying the fluid and for removing small particles arising
from fluid degradation. The paper will additionally provide a brief description of the
design requirements of the steam-turbine electro-hydraulic control system, together
with an explanation of the degradation mechanisms of phosphate esters, the
products of degradation, and their impact on fluid life and performance. An
introduction to the principal factors affecting the efficiency of different ion-
exchange treatments follows, and the paper concludes with a discussion on the
work required before a final resin selection can be made.

Introduction
In the 1950s, an increasing demand for greater stability of the power supply forced
the steam-turbine manufacturers to provide a more efficient turbine control system.
Historically, the function of controlling turbine speed and the rate of load increase
was performed by a mechanical hydraulic control (MHC) system. It utilized a com-
plex arrangement of mechanical devices operating at relatively low pressures.
Unfortunately, this system had a limited control capability.
To provide a faster response to a changing load on the electric grid and to pro-
vide more precise control, an advanced turbine control system was developed for
the majority of new turbines in nuclear-power-generating stations. This new sys-
tem, called electro-hydraulic control (EHC), utilized a complex electronic control
system and a dedicated high pressure (typically between 1000 to 2000 psi) hydraulic
system independent of the turbine lubrication system.
During the initial development stage of EHC systems, more attention was given
by turbine manufacturers to the design of the electronics, considering the hydraulic
portion as a relatively mature design. Consequently, different designs of EHC sys-
tems exist today that have a different impact on system reliability.
Regardless of this new approach, turbine trips occasionally occur resulting in a
total loss of production. According to the Electric Power Research Institute (EPRI),
EHC system failures caused by both electronic and hydraulic component issues are
a significant contributor to such trips [1]. The objective of this paper is to address
some of the critical hydraulic issues affecting the EHC system reliability.

Characterization of Typical Electro-Hydraulic


Systems
The EHC system is a complex design consisting of two major systems: electric/elec-
tronic control systems and the hydraulic fluid system. In addition, it also has an
40 STP 1573 On Fire Resistant Fluids

emergency trip system, valve controllers, and a monitoring system. Although this
system is vital for power generation, it is not considered a safety-related
system. However, many nuclear regulatory agencies include requirements for main-
taining the hydraulic fluid within the industry specification in the station operating
license.
The hydraulic portion of a typical EHC system consists of the steam control
valve system, emergency trip system valves, pumps, accumulators, coolers, filters,
instrumentation, piping, reservoir, and the fluid purification system. Each of these
components has a different impact on the fluid deterioration rate. Usually pumps,
reservoirs, pressure-relief valves, and the piping arrangement create the largest
impact on fluid degradation rate, whereas steam control valves, specifically, the
servo valves and solenoid valves, are the most sensitive components, failures of
which are usually cited by EPRI as the largest category problem in the EHC systems
[1–3]. Servo valves are particularly sensitive to the fluid condition and, therefore,
monitoring the fluid properties is essential to prevent this failure mode [4].
In addition, the EHC system has a few interconnections with other plant sys-
tems, such as the cooling water system for the heat exchangers, compressed air sys-
tems for the air operated valves, and the low voltage electrical distribution system
for powering various motors and solenoid valves. Malfunction of these systems also
influences the reliability of the EHC system.

HYDRAULIC FLUID PUMPS


The pump trains of all EHC systems are designed with either two pumps, each ca-
pable of 100 % capacity, or three pumps, each capable of 50 % capacity. The major-
ity of EHC systems use positive displacement pumps (e.g., vane, piston, screw) and
have parallel fluid delivery systems associated with independent pumps, filters,
pressure-relief valves, and instrumentation. Normally, only one pump line is in
operation with the other in standby mode. The standby pump starts when lower
pressure is detected in the common discharge header.
Until recently, many turbine manufacturers used constant volume pumps. In
such systems, typically 95 % of the fluid is returned to the reservoir over a pressure-
relief valve. Apart from their high energy consumption, such systems create addi-
tional stress on the fluid, which is manifested by a higher fluid degradation rate.
The continuously high flow rates encourage the entrainment of air as a result of tur-
bulence and with poor tank design this can lead to a condition known as micro-
dieseling where compression of the air bubbles in the pump creates very high tem-
peratures and localized degradation. Although adiabatic compression, the most
extreme condition, is unlikely to be seen, the temperatures may still be high enough
for the complete destruction of the fluid molecule [5].
General Electric was the first turbine manufacturer to introduce variable vol-
ume pressure compensated pumps into their EHC systems. These pumps vary the
discharge flow capacity based on system demand. Such a system not only reduces
system power consumption but also provides a more stable discharge pressure and
PHILLIPS ET AL., DOI:10.1520/STP157320130102 41

significantly less stress on the fluid. Currently, this type of pumping arrangement is
the preferred design approach for all new and retrofitted EHC systems.
The major source of pump problems is reported to be contamination of
hydraulic fluids [2]. Other common problems are related to pump cavitation, par-
ticularly because of negative suction, air entrainment, and cold fluids. Excessive
vibration is also cited as another common pump problem [6].

HYDRAULIC FLUID RESERVOIR AND PIPING


A typical EHC fluid reservoir provides a storage, supply, and return location for the
EHC hydraulic fluid. It is a stainless steel tank with internal baffling to increase
fluid residence time and promote release of entrained air. The size of the fluid reser-
voir varies significantly between different turbine manufacturers and depends on
selected fluid/pressure flow design criteria. The reservoir size ranges from approxi-
mately 750 litres (for small Westinghouse EHC systems) to 75|000 litres (for large
Siemens EHC systems). Each reservoir would have a number of cover plates to pro-
vide access for cleaning and inspection.
A small reservoir may promote adiabatic compression particularly if the fluid resi-
dence time is short and is insufficient to release any air bubbles. In most cases, the use
of such reservoirs would relatively quickly result in signs of fluid degradation. However,
there are some advantages. The fluid may respond faster to purification treatment
because of the smaller volume and because the cost of purification media would be sig-
nificantly lower. Also from a maintenance point of view, transferring a smaller amount
of fluid to a temporary storage tank is a relatively easy process.
The larger reservoirs can improve fluid residence time allowing a more effective
separation of entrained air and precipitation of solid particles. However, their use
may delay the appearance of significant fluid deterioration. In addition, any fluid
purification process would be more difficult to perform, time consuming, and more
expensive.
Regardless of reservoir size, fluid level should be maintained close to the upper
limit to maximize its residence time. Because of the inherently hygroscopic nature
of phosphate esters, moisture control in the fluid reservoir is an important factor.
In many cases, air entering the tank passes through a desiccant breather, which is
usually associated with a 5 lm air filter (to reduce contamination by dust). Fan-
assisted extraction may also be used to maintain a constant flow of air across the
fluid surface. Yet other designs pipe dry air into the reservoir. Occasionally, reser-
voirs may also be purged with ambient air. Such a system can be effective if the sur-
rounding air has a relatively low humidity; otherwise, the fluid may become
significantly contaminated.
Active head-space membrane dryers were recently introduced as a cost-effective
source of dry air. This system applies membrane technology using a hollow, fiber-based,
membrane dryer to separate moisture from the compressed air. In addition, this system
also applies a pre-filter for the removal of solid particles and any oily residues from the
air before allowing the dry air to purge the fluid reservoir.
42 STP 1573 On Fire Resistant Fluids

Some reservoirs may have electric heaters. They may burn the fluid layer adja-
cent to the heater tubes if the heat density is too high and there is no fluid circula-
tion. Usually, low-density heaters are applied for this application or the heat is
applied indirectly. Some systems heat the fluid by re-circulating it within a closed
circuit. This would require additional small centrifugal pumps forcing the fluid to
flow through the reservoir.
Another problem may be related to the existence of large gaps, including the
drain openings, in the reservoir baffle plates. These can cause the fluid to bypass the
designed fluid path and shorten the fluid residence time.
The piping in EHC systems is normally made from stainless steel, which dis-
plays good fluid compatibility. However, with poor system design, its use may still
influence fluid degradation. For example, the location and the depth of the return
line in the tank are important factors affecting fluid life. Locating the return line
close to the pump inlet could significantly reduce the fluid residence time and pro-
mote aeration, particularly if it is not submerged at all times in the fluid. Another
factor is the slope of the return line to the reservoir. A greater slope will promote
more turbulence and aeration. This may create an additional problem at the reser-
voir with the air bubble separation process. Close proximity of a fluid line to a
steam pipe, particularly if the latter is not properly insulated, could cause local heat-
ing of the fluid [7].
There are several reports describing problems relating to the use of unsuitable
sealing materials or inadequate maintenance. These can also adversely affect the
fluid condition and reduce the EHC system reliability [8].

STEAM CONTROL VALVES


Steam valve configuration and operation vary widely depending on turbine manu-
facturer and model. The main stop- and reheat valves are used to stop the steam
flow to the high- and low-pressure turbines, respectively. The function of control-
ling the steam flow rate to high pressure turbines is provided by the governor valve,
whereas a similar function for low pressure turbines is done by the intercept valves.
The exact operating sequence and function of these valves during different stages of
turbine operation may vary and is a complex process outside the scope of this
paper.
In addition, turbine manufacturers have different approaches for positioning
turbine control valves (e.g., the governor and interceptor valves). Some of them
applied servo valves with close dimensional tolerances, others applied electrical sol-
enoid valves operated by hydraulic servo motors.
Servo valves receive an electrical signal from the control system causing a
movement of the valve nozzle to position the spool valve. This will set the servo
valve in either the open or closed position. Servo valves also have a bias spring,
which is designed to position the spool valve such that the valve will close on a loss
of electrical signal to the servo coil [1]. Excessive internal leakage is usually caused
PHILLIPS ET AL., DOI:10.1520/STP157320130102 43

by electrochemical erosion as a result of deteriorated fluid condition [4]. Another


frequent problem is a sticky servo-valve condition as a result of the presence of
metal soaps or other degradation products, which overcome the ability of the elec-
tric signal or the bias springs to move the spool.
A solenoid valve is a valve having an electrical coil and a plunger. The solenoid
portion is an electromagnet. When a voltage is applied, a magnetic field is generated
causing the plunger to re-position. An attached spring forces the plunger to return
to its original position when the voltage is removed [1]. The most frequently
reported failure modes are binding or internal leakage and both failures are attrib-
uted to contamination of the hydraulic fluid.

Fire-Resistant Fluids for EHC Systems


High fluid pressures applied in EHC systems increase the risk of a fire arising from
the escape of a jet or spray of hydraulic fluid onto a hot steam pipe. To reduce the
fire risk, turbine manufacturers selected a fire-resistant hydraulic fluid for this
application.
Originally two different fluid types were used: chlorinated aromatic hydrocar-
bons and triaryl phosphate esters. The use of the former was discontinued in the
1970 s because of servo-valve erosion problems and environmental concerns [9]. As
a result, phosphate ester fire-resistant fluids have been used for over 50 years as tur-
bine control fluids at power stations. Whereas they have been successful in prevent-
ing fires, their use has not been trouble free.
Triaryl phosphate esters were initially introduced as fire-resistant hydraulic flu-
ids towards the end of World War II following a series of fires and explosions in
military hardware, particularly aircraft, and their use in industrial applications
slowly followed. Their initial use in power generation probably occurred in Ger-
many as a steam-turbine main bearing lubricant (and hydraulic fluid) following
major turbine fires. A successful trial was carried out in 1942 in a 6 mW industrial
steam turbine, and a report was made to the German Power Station Association in
1943 of over 6000 operating hours without problems [10].
Phosphate esters are also known as esters of phosphoric acid and can be con-
sidered as having the following general structure (Fig. 1).
Although trialkyl phosphates are available, only triaryl phosphates are used in
power generation. In these fluids, R1–R3 are phenyl or substituted phenyl (e.g.,
methylphenyl, also known as cresyl or tolyl), dimethylphenyl (better known as
xylyl), isopropylphenyl, or tertiarylbutylphenyl groups.
The raw materials (phenol and substituted phenols) available for the manufac-
ture of triaryl phosphates have changed over the years, and, as a result, different
types of triaryl phosphate have become commercially available. Most of these phos-
phates have at some time been used in EHC systems, but in the last 20 years the
industry has focused on three main types: isopropylphenyl phosphates, tertiarybu-
tylphenyl phosphates, and xylyl phosphates. Although all provide a high level of fire
44 STP 1573 On Fire Resistant Fluids

FIG. 1 The generic structure of neutral phosphate esters.

resistance, their physical and chemical properties do vary, and, depending on oper-
ating conditions, this can influence their selection. These products have, therefore,
been used either as 100 % fluids or in blends with each other.

Phosphate Ester Degradation Processes and the


Products
of Degradation
Phosphate esters can degrade in several ways. Whereas this discussion used to be
limited to a simple study of hydrolysis (the reaction with water) and oxidation
under laboratory test conditions, there is now an awareness that conditions in serv-
ice can be considerably more severe than those simulated in the laboratory. For
example, the very high temperatures found as a result of micro-dieseling (rapid
compression of air bubbles) and static discharge, considerably increase the com-
plexity of reactions taking place in the fluid and therefore the range of degradation
products that can arise during decomposition. This section looks at the different
degradation mechanisms, the products of degradation and the sensitivity of the dif-
ferent types of phosphate to the main modes of degradation. More detailed infor-
mation on the effect of degradation on phosphates is available in Ref 11.

HYDROLYTIC STABILITY
In the presence of small amounts of water, phosphate esters tend to hydrolyze [i.e.,
break down into their constituent acids and alcohols (or, in the case of aryl phos-
phates, acids and phenols)]. It is the most common form of degradation of phos-
phates and the one that most frequently dictates their service life.
The hydrolysis reaction is not rapid at ambient temperatures but accelerates
with increasing temperature and is catalyzed by the presence of strong acids and
PHILLIPS ET AL., DOI:10.1520/STP157320130102 45

FIG. 2 The hydrolytic stability of different phosphate esters showing the change in rate
with time.

some metals. As the products of hydrolysis are themselves strong acids, the process
is said to be “autocatalytic.” Strong bases can also catalyze the reaction. The rate of
change of acidity increases with time as shown in Fig. 2 [11], which compares the
stability of different types of fluid in the standard hydrolytic stability test (ASTM
D2619-09 [12]). This is, however, an accelerated test because of the high water con-
tent and high surface area of the copper catalyst relative to the volume of fluid used.
The hydrolysis of triaryl phosphates (e.g., triphenyl phosphate) takes place as
indicated in Fig. 3 [13]. The replacement of successive aryl groups with -OH groups
becomes increasingly difficult and the last step is not normally achieved without
using very severe conditions, such as boiling with dilute acid or at very high temper-
atures (e.g., under micro-dieseling conditions). As a result, under normal operating
conditions, it is extremely unlikely that phosphoric acid would be generated.
The reaction scheme shown in Fig. 3 shows the strong acids (pH < 4)—also
known collectively as partial phosphates—that are produced. Although these are of
greatest importance in view of their reactivity, weak acids (pH 6–7) such as phenols,
are also formed in this process. Whereas they are not thought to have a major effect
on the performance of the fluid, they can adversely affect foaming and volume resis-
tivity, the latter being used to assess the potential for servo-valve erosion. They are
also intermediates in the formation of polyphosphates arising from the condensa-
tion of two or more mono- or dihydrogen phosphates and in the presence of metal
soaps can form metal phosphates.
The presence of strong acids can have different effects on fluid and system per-
formance. Some of these are beneficial, whereas others can cause operational prob-
lems. The acids may, for example, inhibit the corrosion of ferrous metal surfaces
but promote corrosion of nonferrous components; they may adversely affect the
electrical properties of the fluid but assist in reducing wear. In general, the
46 STP 1573 On Fire Resistant Fluids

FIG. 3 The hydrolysis of triphenyl phosphate.

uncontrolled generation of acidic products is harmful to the life and performance


of phosphate esters, and in certain critical applications (e.g., in power generation),
it is customary to remove them as they are produced by circulating the fluid
through an adsorbent solid thereby prolonging fluid life and this will be discussed
in greater detail later in this paper. However, strong acids are also chemically reac-
tive and can form metal soaps or salts, for example, with components of the fuller’s
earth or activated alumina adsorption media and these are subsequently dissolved
or dispersed in the fluid. As the soaps eventually precipitate (e.g., in servo valves)
causing sticking and also promoting foaming and air retention, their presence is
highly undesirable and is a major reason why the industry has moved to the use of
ion-exchange resins.
The importance of acidity in the fluid degradation cycle is shown in Fig. 4 [13]
and underlines the need to reduce the acid level and find alternatives to the fuller’s
earth and alumina adsorbents.
Whereas, utilizing the most efficient acid adsorbent is important, the use of a
phosphate with superior hydrolytic stability is also beneficial and such materials are
currently used for the main bearing lubrication of steam turbines where no in situ
conditioning or purification is provided.

OXIDATION STABILITY
Good oxidation stability is an essential property for hydraulic fluids that are
exposed to “high” temperatures. Oxidative degradation results in the production of
a range of acids (both strong and weak) and from very low to high molecular
PHILLIPS ET AL., DOI:10.1520/STP157320130102 47

FIG. 4 The fluid degradation cycle.

weight; low molecular weight hydrocarbons and their oxidates that may plate out as
varnish and higher molecular weight hydrocarbons or polyphosphates present as
sludge that can increase fluid viscosity and reduce resistivity while blocking filters
and causing valve sticking. At very high temperatures, extremes of oxidants, for
example, hydrogen and carbon (see below) will be formed. However, in most hy-
draulic systems, fluids are intermittently subjected to temperatures up to
150 C–200 C as a result of being compressed in pumps, bearings, relief valves, and
restrictors [14], and sometimes from tank heating. In addition, fluid can be exposed
to heating from external sources such as a steam line, hot/molten metal, or a weld-
ing torch located close to the hydraulic line. The difference between internal “hot
spots” and external heat sources is that the former occur when the fluid is in circu-
lation (except for the fluid being heated up in the tank), whereas an external heat
source can also be present when the fluid is static. If the fluid is circulating, oxida-
tion will normally occur because of the presence of air. However, if the fluid is
static, oxidation may occur until the oxygen is consumed, after which pyrolysis (or
thermal degradation) may take place.
Oxidation is a free radical process occurring as a result of energy applied in the
form of heat. The first step normally involves an attack on the alkyl substituent on
the ring. The isopropyl group is particularly vulnerable because of the labile second-
ary hydrogen atom; the methyl group is more stable and the tertiary butyl group is
very resistant to attack. Unsubstituted phenyl groups are also extremely stable and
for both this and the tertiarybutylphenyl phosphate, the initial degradation arises
through thermal breakdown (fission) of the P–O bond.
The isopropylphenyl phosphates and TCP/TXP are responsive to the addition
of stabilizers, and in service can contain inhibitors that improve performance and
significantly extend life. In contrast, the tertiarybutylphenyl phosphates are not as
responsive to classical antioxidants but have such good oxidation stability that this
is rarely a disadvantage.
48 STP 1573 On Fire Resistant Fluids

TABLE 1 Coking tendency data for triaryl phosphates: Test method: Federal test method standard
VV-L-791 C, method 3462 [16].

Deposit Formation (mg) At

Phosphate Ester 300 C 316 C 325 C

TXP (uninhibited) 460 1820 1750


TXP (inhibited) 43 1006 1420
TBPP/32 (uninhibited) 3 16 6
TBPP/46 (uninhibited) 3 4 2
IPPP/46 (inhibited) 4 – –
Mineral gas turbine
lubricants
ISO VG 32 25 170 130
ISO VG 46 49 185 227

One additional aspect of oxidation stability that is of concern in some applica-


tions is the deposit-forming tendency at high temperatures. This property, also
known as “coking,” occurs when a thin film of fluid is heated on a metal surface
while exposed to air. Depending on the stability of the product, the deposit can vary
from soft and carbonaceous to a hard, brittle layer, or to a lacquer. The formation
of such deposits can occur on heater and valve surfaces and can reduce component
efficiency. In situ conditioning, however, has been shown to control the deposit for-
mation [15]. Another form of deposit, commonly known as “varnish” occurs when
fluid degradation products precipitate onto the surface of metal components and
then harden under the influence of heat. It can be difficult to differentiate between
the two forms of deposit without complex analytical investigations. Table 1 [16]
shows the variation in coking tendencies for a range of triaryl phosphates and com-
mercial mineral gas turbine oils.
Mention was earlier made of a phenomenon called micro-dieseling. This
involves the compression of air bubbles in the pump, which have not been released
during the retention time in the tank. Unless the bubbles dissolve readily (which
depends on their size), dieseling is accompanied by a significant increase in the tem-
perature of the wall of the bubble. When no heat is lost from the bubble as it is
compressed, the process is known as adiabatic compression, and temperatures
within the air bubble as high as 1000 C have been calculated for pressures of about
140 bar [17]. However, in practice, some heat is lost to the surrounding fluid and
temperatures are unlikely to be so high (unless the pressure is increased still fur-
ther). Certainly, evidence exists for temperatures around 800 C–850 C [5], but
even this temperature greatly exceeds the thermal stability of the molecule and it
will cause fragmentation with the lower molecular weight species reacting with oxy-
gen or polymerizing. In mineral oil, these high temperatures have been shown to
PHILLIPS ET AL., DOI:10.1520/STP157320130102 49

FIG. 5 Some possible high-temperature degradation mechanisms.

cause combustion (a discharge) within the air bubble but to date there have not
been any reports of this occurring with phosphates.
In addition to the decomposition products found at lower temperatures, the
very high temperatures (found on, and adjacent to, the bubble wall) of this process
also mean that other reactions can take place involving nitrogen rather than oxy-
gen. For example, fixation of nitrogen has been reported as well as the formation
of nitrogen oxides [17]. These are the “raw materials” for a range of additional
degradation products, several of which have been identified in the analysis of
phosphate samples taken from systems where dieseling is occurring. Some of these
degradation products are indicated in Fig. 5 [5], whereas the main gaseous degra-
dation products have been identified by dissolved gas analysis and are listed in
Table 2 [5].
One of the features of dieseling is that the fluid rapidly darkens with the pro-
duction of copious amounts of sub-micron particles, most of which are carbon.
Although these particles are too small (and, normally, too soft) to cause significant
wear, they do have an adverse effect on other fluid properties such as foaming, air
release, and volume resistivity. They can also blind filters.
The effects of static discharge on the composition of degradation products has
not yet been established for phosphate esters but is anticipated to be similar to the
effects of micro-dieseling.

THERMAL STABILITY AND PYROLYSIS


The thermal stability of a fluid can provide an approximate guide to its upper oper-
ating temperature, but then only in terms of its ability to withstand breakdown in
50 STP 1573 On Fire Resistant Fluids

TABLE 2 The results of dissolved gas analysis by ASTM D831 on samples taken from EHC systems
where dieseling is taking place.

Turbine A Turbine B Turbine C Turbine D Fresh fluid


Gas (ll/l) (ll/l) (ll/l) (ll/l) (ll/l)

Oxygen 400 7587 11 811 16 155 200 095


Nitrogen 44 200 38 868 38 493 38 909 37 120
Carbon 10 100 195 582 937 ND
monoxide
Carbon dioxide 13 950 801 3366 4594 651
Hydrogen 1330 55 126 175 ND
Acetylene 1800 128 626 963 ND
Ethane 230 12 55 92 ND
Ethylene 860 64 225 307 ND
Methane 1900 38 213 333 ND

the absence of air or oxygen (pyrolysis)—a situation that is rarely found in practice.
In reality, the presence of a small amount of dissolved oxygen will result in degrada-
tion at lower temperatures—particularly in the presence of metals—and apparent
changes in the physical/chemical properties of the fluid may be because of oxidation
rather than pure thermal breakdown.
Several studies have been carried out into the products of pyrolysis and com-
bustion of phosphate esters. Lhomme et al. [18] examined the degradation products
under helium of trimethyl, triethyl, and triphenyl phosphate, and, as a result, pro-
posed a general pyrolysis scheme (Fig. 6). Depending on the phosphate structure
and temperature, different degradation pathways were followed. At “low” tempera-
tures, phosphates followed reaction (a) as a result of the cleavage of the –C–O bond
with the production of olefins, monohydrogen phosphates, dihydrogen phosphates,
etc. At “higher” temperatures, path (b) was followed, involving the breakage of the
–P–O bond, while at very high temperatures, the phosphate residue forms phos-
phorus pentoxide (probably after passing through an intermediate phase involving
the formation of polyphosphates and pyrophosphates). The pentoxide rapidly
reacts with moisture in the atmosphere, first to form metaphosphoric acid and then
phosphoric acid itself.
With triphenyl phosphate, reaction (b) predominated with initial decomposi-
tion not occurring until >600 C and not being complete at 1000 C. It seems highly
probable that where there is substitution on the aromatic ring, degradation would
initially follow reaction (a), and then (b) when unsubstituted rings were formed. It
is interesting that both processes produce carbon, a significant contaminant when
dieseling occurs.
Pyrolysis studies of IPPP/46 under a helium atmosphere have also been carried
out at temperatures between 500 C and 1000 C (Table 3) [16]. Measurements were
PHILLIPS ET AL., DOI:10.1520/STP157320130102 51

FIG. 6 General pyrolysis scheme for phosphates.

made of (1) the amounts of carbon monoxide and dioxide formed using non-
dispersive infrared analysis, (2) the organic volatiles, which were identified using
gas chromatography/mass spectrometry, and (3) the amount of phosphorus pent-
oxide generated, which was collected in aqueous potassium hydroxide and deter-
mined as orthophosphate by ion chromatography. For comparison, the same
product was also examined for the production of carbon dioxide and phosphorus
pentoxide under combustion conditions by passing air over the sample. Under
combustion or oxidative decomposition, the product shows significant degradation
at the latter temperature, but the values for phosphorus pentoxide content were
lower than might be expected.
The three main degradation processes are thus responsible for the generation
of a wide variety of degradation products. If these are to be minimized, the fluid
should be kept clean and dry, and as close to ambient temperature as possible.
52 STP 1573 On Fire Resistant Fluids

TABLE 3 Development of carbon oxides and phosphorus pentoxide under pyrolysis and combus-
tion conditions for an IPPP 46 phosphate.

Temperature ( C)

500 700 1000

Yield of
Degradation Pyrolysis Combustion Pyrolysis Combustion Pyrolysis Combustion
Products
(%)

Carbon 0.63 – 0.56 0.1 1.1 11.23


monoxide
Carbon <0.001 – <0.001 0 <0.001 49.63
dioxide
Phosphorus <0.01 – <0.01 0.57 <0.01 1.42
pentoxide

However, attempting to reduce the level of degradation will be difficult to achieve if


the system design encourages oxidation to take place or allows moisture ingress, or
if the fluid has poor stability.

Purification Treatment
The principal behind all conditioning systems is that the rate of acid removal must
be greater than the rate of production. If not, the control of the fluid degradation is
lost, the acidity will increase quickly and fluid life will be considerably shortened.
However, purification is not just about the removal of acid but also about keeping
the fluid dry and clean.

HISTORICAL BACKGROUND
In 1966, Schober [19] reported on the successful continuous treatment of phosphate
esters by fuller’s earth (an Attapulgus clay based on aluminosilicates) on a bypass
system to the main fluid reservoir of a steam turbine. Continuous (rather than
intermittent) treatment was found to have advantages in extending the life of the
fluid and keeping the fluid dry as a result of water adsorption by the earth. In addi-
tion, continuous treatment used less earth. The possibility of interaction between
the earth and the fluid degradation product was discussed and a pyrophosphate (or
a mono-valent metal salt thereof) was extracted from used earth. The inference was
that the aluminosilicate in some way catalyzed the formation of the pyrophosphate
in addition to the conventional removal of acid.
The use of fuller’s earth to remove “active chemical species” from phosphate
esters that were causing servo-valve erosion was reported by Wolfe and Cohen [9].
They too attempted to identify compounds that were removed from the erosive
fluid. In addition to finding chloride present, the magnesium salt of di-m-cresyl
PHILLIPS ET AL., DOI:10.1520/STP157320130102 53

phosphoric acid (Fig. 9) was also identified and thought to be a reaction product of
the fuller’s earth and fluid degradation products.
The first reports of deposits when using fuller’s earth on turbine control sys-
tems arose in the late 1970 s, primarily in France and Germany. In France, a survey
in 1978 [20] suggested that not all power stations at that time were fitted with fluid
purification units, and those that were installed were undersized, i.e., had insuffi-
cient fullers earth to control acidity generation. There was also a general lack of
appreciation of the importance in controlling acidity generation. As a result, fluid
acid values in problem stations had risen to 5–10 mg KOH/g. Where earth treat-
ment was used, such high acidities were associated with deposits, which were found
to contain calcium and magnesium salts.
As an alternative adsorbent, activated alumina was evaluated and found to offer
fewer problems in this respect. Although sodium could be extracted in significant
amounts, no deposits were observed. It was, however, reported that alumina was
less effective than fuller’s earth in improving the resistivity, corrosion, and oxida-
tion performance of fresh fluid [21].
In Germany, a similar report of system deposits was made by Grupp in 1971
[22]. Analysis of the gelatinous materials revealed the presence of a variety of metal
soaps, some of which arose as a result of system corrosion and others, e.g., calcium/
magnesium, arose from the fuller’s earth. To clarify the origin of the calcium and
magnesium, the composition of earths from different suppliers was investigated
and the presence of calcite (CaCO3) and dolomite (CaCO3MgCO3) were identified
as the sources. It was proposed that the mono- and di-hydrogenphosphate degrada-
tion products reacted chemically with the calcium and magnesium carbonates to
form the metal soaps. Depending on the degree of hydrolysis, the molecular weight
of the salts varied, with the possibility of the production of polymeric materials of
high molecular weight.
As a result of this investigation, the formation of deposits, especially during
shutdowns or on other occasions when the fluid was allowed to cool, was forecast if
the total content of calcium and magnesium exceeded 30 ppm. It was also recom-
mended to use an earth free of carbonate content. However, the manufacturers sub-
sequently indicated that, as fullers earth was a natural product, this aspect could not
be guaranteed. As a result, problems of this type continued, and in 1992 Grupp pro-
vided additional information on the sensitivity of clays to the presence of both
water and acid [23].
In addition to the problems associated with metal soap formation, several other
operating difficulties are associated with fuller’s earth, in particular, the need to dry
the earth before use as the presence of water reduces its efficiency, and also its occa-
sional “lack of activity” (there have been a number of cases where fuller’s earth
replacement has failed to have any significant impact on acid removal). Normally,
the problem arises at “high” acid levels. It could be caused by a high water content
or a lower rate of acid removal in comparison with the rate of formation, but has
not been investigated in any detail.
54 STP 1573 On Fire Resistant Fluids

As an alternative to fuller’s earth, some utilities converted to using activated


alumina, a synthetic product with a known and guaranteed composition. In a few
cases, the change was made because alumina seemed to be effective when fuller’s
earth failed to reduce acidity levels, and utilities would use alumina to reduce the
acidity to an acceptable level before replacing it with fuller’s earth. A reluctance to
use alumina on a continuous basis can be attributed to a higher price and perhaps
also to the fact that the commercial aluminas had a wider particle size distribution
than the earths. There was, therefore, a greater possibility of “fines” escaping from
the cartridge into the system and, being hard particles, promoting wear. It was nec-
essary—perhaps even more important—to dry the alumina before use as this solid
absorbs more moisture than the fuller’s earth.
Most of the experience quoted above has focused on the presence of deposits
and their avoidance by using dry fluid or by changing or modifying the composi-
tion of the solid adsorbent. In contrast, little attention was paid initially to control-
ling the acidity. Schober [19] and Tersiguel-Alcover [20] suggested that the
maximum level in service should be 1 mg KOH/g. However, it became apparent
that even this level was too high to avoid deposit formation and turbine builders’
recommendations for limits on in-service fluid began to drop, initially to 0.5 mg
KOH/g and then to 0.3 mg KOH/g. A report [24] published in 1979 on the effec-
tiveness of different types of adsorbents advised maintaining an acidity of <0.2 mg
KOH/g, and subsequently even lower values were suggested, as long as the fluids
were treated with adsorbent solids that contained extractable metals. A low acidity
has, of course, the advantage of reducing the sensitivity of the fluid to hydrolysis as
the process is acid-catalyzed, and, whereas the maintenance of low acid levels has
an obvious effect on fluid monitoring and treatment requirements, the slightly
increased analytical and labor costs are normally small in comparison with the cost
of a replacement fluid charge.
Molecular sieves have been proposed for conditioning phosphate esters [25],
preferably in conjunction with a coalescer/separator to remove water. In the writers’
experience, however, molecular sieves do not remove partial phosphates from solu-
tion and can themselves promote metal soap formation at relatively high acidity
levels. This technique has not gained widespread acceptance.
The operating problems associated with the use of fuller’s earth and alumina
have led users, turbine builders and fluid suppliers to investigate other potential
conditioning agents, and two separate technologies have arisen in the last 20 years.
The first is the development of a “synthetic” fuller’s earth, also known as a Y zeolite,
in combination with a purer form of activated alumina [26]. The material is very
effective at adsorbing acids but can only be used with fresh fluid or fluid that has
not been significantly degraded. When acidity levels increase, it still appears possi-
ble to form metal soaps, although the tendency is significantly less than with
“natural” fuller’s earth. As the solid releases water during acid adsorption, it is nec-
essary to use a water-absorbing filter immediately downstream of the solid. Some
concern has also been reported regarding slow improvement of fluid resistivity and
PHILLIPS ET AL., DOI:10.1520/STP157320130102 55

the possibility of hard particles of alumina escaping into the system. The other de-
velopment has been the introduction of ion-exchange (IX) resins, and these are dis-
cussed more fully in the following section.

ION-EXCHANGE (IX) SYSTEMS


The possibility of using ion-exchange resins was originally investigated by Wolfe
and Whitehead [27] in 1977 in experiments with strong base anionic resins.
Although effective in reducing acidity, this technique was rejected because of price,
the need to use “wet” resins, as well as the fouling of resin beads with gels and sili-
cone anti-foams. Wolfe and Whitehead, however, only looked at acidity reduction,
rather than removal of metal soaps, and also used degraded fluid with an initial
acidity of 2–3 mg KOH/g, far in excess of what is typical (or permissible) today.
The next step forward was taken at Gösgen nuclear power station (Siemens
990 MW nuclear steam turbine) in Switzerland in 1983 [28]. As a result of the fail-
ure of fuller’s earth to control acidity generation in a “synthetic” fluid (and because
of associated foaming problems and deposit formation on valves), the possibility of
replacing the earth was investigated. Laboratory tests indicated that a weak base an-
ionic resin (WBA) in the hydroxide form could quickly reduce the acidity, whereas
a strong acid cationic resin (SAC) would, surprisingly, reduce the metal soap con-
tent. A decision was made to use these resins in the existing bypass loop in conjunc-
tion with a molecular sieve to remove the water released as the resins are normally
supplied “wet,” and release water in use. To reduce the amount of water released by
the resins, they were dried at 80 C to a level of 5 % before use as a mixture in the
existing filter housing. Over a period of about 1–2 months, the properties of the
fluid returned to close to those of a new fluid (Fig. 7). When the metal soaps had
been reduced to a very low level, the treatment reverted to the use of a WBA resin

FIG. 7 Change in fluid properties at Gösgen Power Station during early stages of resin
treatment.
56 STP 1573 On Fire Resistant Fluids

alone. The fluid continued to operate without problems and in good condition until
it was replaced in 2010 after 31 years, not because the product had deteriorated but
because the fluid was being withdrawn. This performance was achieved using inter-
mittent treatment (approximately 2 days, 2–3 times a year) and the resin charge
was replaced every 3 years [29].
The success of the Gösgen experiment resulted in wider use of the resins in
Switzerland and Germany, where experience was also very positive [30] and
prompted a more systematic investigation into their potential [31].
In Canada, the first country outside Europe to adopt the new technology, it was
reported that by 2001 a major gas pipeline company that operated about 60 turbo-
compressors on phosphate esters, had accumulated 6  106 operating hours on ion-
exchange resins. As a result, considerable savings were being made, mainly on fluid
replacement costs [32].
Also in Canada, Ontario Power Generation introduced resin treatment in 1993
and the subsequent operating experience has been well documented [33–35],
whereas information on its use in the United States is reported in reference [36].
Use of this technology has now spread to many utilities in Europe, the United
States, and parts of the Far East. Some changes have, however, been made since the
initial trial. For example, where cationic and anionic resins are used, they are now
mainly contained in cartridges and kept separate from one another.
Whereas, most users initially tried to use wet resins (typically 50 %–65 %
water), not all were successful. The need to dry the fluid after ion-exchange treat-
ment (perhaps using vacuum dehydration) to reduce the amount of water present
encouraged the development and use of “dry” resins, normally containing less than
10 % water. It was also realized that whereas WBA resins were useful in many
applications, there were contaminants that responded better to strong base anionic
(SBA) or strong acid cationic (SAC) resins. Various resin mixtures have therefore
been proposed depending on the fluid composition.

RESIN COMPOSITION AND STRUCTURE


Chemically, ion-exchange resins are based on a divinylbenzene or polystyrene core
onto which different functional groups have been introduced. These functional
groups can react with other chemicals leading either to an exchange of ions or to
the adsorption of the chemical onto the resin surface. As a result, unlike fuller’s
earth or Selexsorb, they can remove different degradation products or contaminants
from solution. The function of commercially available resin types is described in
Table 4 [37].
In general, there are three different types of resin structures where the polymer
matrix differs:
• Gel type. These resins are translucent and the pore structure depends on the
degree of crosslinking (or amount of divinylbenzene—DVB) which is usually
in the region of 8 %–10 %. A potential disadvantage of this type is the large
change in volume as the ionic form changes and the small pore size results in
PHILLIPS ET AL., DOI:10.1520/STP157320130102 57

TABLE 4 Function of commercially available resin types.

Resin Type Benefits Disadvantages

Weak base anion Removes strong acids and metal Weak acids remain
soaps/salts by adsorption

Strong base anion Removes weak and strong acids and Less thermally stable
also will split neutral salts
Functions across full pH range More prone to fouling than WBA
types

Weak acid cationic Will remove neutral soaps Releases acid into the fluid
Less thermally stable
Large volume changes on going from
Hþ to Mþ form

Strong acid cationic Will split neutral soaps and remove Releases acid into the fluid
metals Requires use of chloride-free version
of resin

Chelating Removes acid rapidly May release sodium into the fluid with
possible adverse effect on soap
precipitation and foaming properties

a longer time to reach exchange equilibrium. A gel resin was one of the resins
first used by Göesgen power station for removing the metal soaps by
adsorption.
• Macroporous resins are opaque and contain as much as 20 % DVB (i.e., much
more cross-linking than gel types and are processed to produce much larger
pores than are found in gel types). Because of the higher crosslinking, these
resins are less sensitive to the effects of drying (i.e., show less shrinkage) and
react faster. They also have better strength than gel resins. A particular form
of macroporous (or macroreticular) resin is the so-called chelating resin. This
type of resin has a functional group that has a greater selectivity for certain
metal ions. For example, the amino-phosphonic group will preferentially
remove calcium or magnesium if there is a mixture of ions present.
• Macronet resins have an even higher internal surface area and are rigid (i.e.,
they do not shrink significantly). The exchange capacity is, however, much
reduced. This may not be critical if the mode of action is by adsorption rather
than by ion exchange.

OPERATING EXPERIENCE OF IX TREATMENT AT CANDU STATIONS


As indicated earlier, one of the main differences between the initial application of
the IX resin and those currently commercially available is that originally the resin
was used in bulk. Today, almost all resin use is in cartridge form and their designs
vary, some of them having axial, others radial, flow. Unless the flow rate through
the medium is very slow, axial flow is normally preferred. The resin type is not
58 STP 1573 On Fire Resistant Fluids

FIG. 8 Magnified images (40X) of typical new WBA (left) and two different used resins
(middle, right).

usually indicated, although it appears that most suppliers offer WBA resins. How-
ever, some suppliers also provide a combination of WBA and SAC resin or SBA
and SAC.
Despite previous success with ion-exchange treatment, not all applications have
demonstrated similar benefits. A study was initiated by the CANDU Owner’s
Group to identify and understand factors that may affect the efficiency of IX treat-
ment. As part of this investigation, the authors found that, all resins currently uti-
lized at CANDU stations (both new and used), showed a spherical bead structure.
Most of the spent resins were coated with insolubles to various extents depending
on the EHC system design and the primary degradation mechanism. Water wash-
ing is not always able to remove the coating. Figure 8 [38] shows magnified images
of typical new and used resins from EHC purification systems at CANDU stations.
Usually, in-service fire-resistant fluids are dark, indicating the presence of large
amount of insolubles. Membrane patch colorimetry tests (ASTM D7843-12) carried
out on used fluid samples gave results in the range of 66 to 77 CIE dE), and
although typical ISO Cleanliness Codes were 17/2015/12 or better, significant
amounts of insolubles, particularly in the range of 0.45 to 4 lm, were found.
In stations with a predominately dieseling problem, the control of fluid acidity
is difficult. This is a result of a combination of the impact of system design, inad-
equate acid removal capability of the IX purification system and poor maintenance.
The acid number trends in four different turbines are presented in Fig. 9. These
results demonstrate the difficulty in maintaining fluid acidity below the recom-
mended limit of 0.2 mg KOH/g.
Currently, different resin treatment systems for EHC fire-resistant fluids are
commercially available. However, if end users do not identify the primary degrada-
tion mechanisms and the nature of the fluid contaminants in their EHC systems,
the selected IX system may not provide the expected results. Because of limited
publications on the application of IX treatment, some users perform their own
studies to obtain more details on available options for this treatment.
PHILLIPS ET AL., DOI:10.1520/STP157320130102 59

FIG. 9 Acid number trends in ABB turbines (with permission from Ontario power
generation).

EVALUATION OF DIFFERENT RESINS


A used fluid based on a mixture of isopropylphenyl phosphate and trixylyl phos-
phate fire-resistant phosphate ester was obtained from an EHC system without
servo valves, and was chosen to investigate the ability of different resin types to
reduce fluid acidity. The condition of the fluid is presented in Table 5 [38].
Ten commercially available and three developmental ion-exchange
resins provided by seven different manufacturers were selected for testing (see Table
6). Two resins were received in dry form and all others were dried in the lab by
heating them at 60 C under vacuum until constant weight was obtained.
Nine resins marked from A to F and H, L, and M were tested separately. In
addition, four combinations of resins were also evaluated. These were: G and B, G
and I, G and J, and G and K. They were tested using 3:2 cation to anion ratios.
The test was carried out by passing 3.8 l of fire-resistant fluid through a glass
column containing 200 ml of IX resin in a recirculation mode for 168 h. The
arrangement is shown in Fig. 10. It includes a condenser, a moisture guard to protect
against ingress of atmospheric moisture, and a sampling port. The temperature was
maintained at 40 6 4 C. The flow rate was maintained at 17 ml/min using a peri-
staltic pump. This meant that it took approximately 4 h for a single pass of the
entire fluid through the column. A sample of 50 ml was removed after 4, 24, 48,
120, 144, and 168 h for testing.
Test results are presented in Fig. 11 [38].
Users operating EHC systems with servo valves also need to control volume re-
sistivity to prevent electrochemical erosion of servo-valve spools. Because in the
60 STP 1573 On Fire Resistant Fluids

TABLE 5 Analytical results of used FRF fluid before testing.

OPG Actual
Analysis Units Specification Results

Acid number mg KOH/g <0.2 0.26


Volume resistivity at GX-cm >4 3.84
20 C
Moisture content mg/kg <800 616
Particle count in range 8646
5–10 lm
Particle count in range 3075
10–25 lm
Particle count in range 160
25–50 lm
Particle count in range 19
50–100 lm
Particle count in range 0
>100 lm
ISO 4 lm cleanliness 17 16
code
ISO 6 lm cleanliness 15 14
code
ISO 14 lm cleanliness 12 10
code
ASTM color ASTM scale <5 8
Air release Minutes <8 7.2
Varnish tendency (MPS) CIE dE Not specified 95.1
Foam tendency ml <400 440
Foam stability ml <10 <10
Weight of deposit after g Not specified 0.0135
filtration through
5 -lm filter
Weight of deposit after g Not specified 0.0815
filtration through
0.45 -lm filter
Concentration of Al ppm Not specified 3
Concentration of Na ppm <30 5
Concentration of Ba ppm Not specified 4
Concentration of Li ppm Not specified 16

Note: Concentrations of Cr, Cu, Fe, Pb, Ni, Sn, B, Zn, P, Mg, Ca, Mo, Ag, Si, Ti, and V were below
1 ppm.

past there were some concerns regarding the ability of the ion-exchange resin to
control this property, fluid samples obtained during the IX treatment test were also
tested for volume resistivity. Results are presented in Fig. 12 [38]. It should be noted
that these results were obtained on fluids with a high insoluble content and that, as
PHILLIPS ET AL., DOI:10.1520/STP157320130102 61

TABLE 6 Characterizations of IX resins received from suppliers.

Supplier Resin Resin Type Density (g/ml) Water Retention (%)

Supplier I A WBA 0.6191 49.3


(macroporous)
Supplier I B WBA 0.5749 52
(macroporous)
Supplier I C WBA 0.6503 54.9
(macroporous)
Supplier II D WBA 0.4374 4.4
(macroporous)
Supplier III E SBA (gel type) 0.68 60–65
Supplier III F WAC 1.1 55–60
(macroporous)
Supplier III G SAC 0.74 56–60
(macroporous)
Supplier III H WBA 0.62 54–60
(macroporous)
Supplier IV I SBA (gel type) 0.64 55–65
Supplier V J SBA (gel type) 0.69 54–60
Supplier VI K SBA 0.425 61.01
(macroporous)
Supplier VI L SBA 0.6275 61.51
(macroporous)
Supplier VII M SBA 0.4908 1.8
(macroporous)

Note: WBA, weak base anion; WAC, weak acid cation; SBA, strong base anion; SAC, strong acid
cation.

yet, the effect of the insolubles on the efficiency of acid removal and increase in re-
sistivity have not been clarified. However, it is probable that the activity of the resin
will be adversely affected by surface deposition.
To support selection of the best resin type for the IX treatment additional quali-
fication tests were performed. These tests involved assessment of crush strength
(friability), the percentage of perfect whole beads (PWB), the content of fines and
amount of extractable material (inorganic and organic).
The results of the resin qualification tests can be summarized as follows:
• Crush strength of the six resins listed above varied in the order: D (dry) > H
(dry) > J (dry) > C (dry) > G (dry) > M (dry). Resin qualification tests were
not conducted on “I” dry because it had listed properties that were similar to
“J” dry. M dry resin failed to meet the 3 N/bead crush strength criteria typi-
cally expected of nuclear grade resins used at CANDU stations; hence, addi-
tional qualification tests were not carried out on this resin. All of the other
resins met or exceeded the crush strength criteria.
62 STP 1573 On Fire Resistant Fluids

FIG. 10 Experimental setup for IX column treatment.

FIG. 11 Acid number after treatment by different resins.


PHILLIPS ET AL., DOI:10.1520/STP157320130102 63

FIG. 12 Volume resistivity at 20 C after treatment by different resins.

• All of the resins that passed the crush strength criteria also showed good per-
fect whole bead (PWB) values in the range 96–99.5 % and very low fines
(<0.1 %) content.
• The water extractable residue from the various resins varied in the order: G
(dry) > J (dry) > C (dry) > D (dry) ¼ H (dry).
Taking into account the ability of the resin to reduce fluid acidity and to
improve volume resistivity as well as displaying good crush strength, a high per-
centage of perfect whole beads with a low content of fines and extractables, the pre-
ferred resin for EHC systems with servo valves is the WBA macroporous resin, H
(dry), followed by another WBA macroporous resin, D (dry). However, it was appa-
rent that not all WBA resins have the same ability to control fluid acidity and resis-
tivity as well as possessing good physical/chemical characteristics.
To compare the difference in performance of the most active resins in this
investigation with a resin already in use, two additional plots were prepared. Figure
13 compares the effect on acid number reduction after treatment by two resins cur-
rently used in EHC systems in Canada (A in wet and dry form) and the perform-
ance of the two new recommended resins (D and H in dry form).
Figure 14 compares the similar effect on volume resistivity reduction between
the current (A, wet and dry) and the new resins (D and H in dry form).

CONTROL OF FLUID MOISTURE


A separate test was also carried out to evaluate three different techniques for re-
moval of moisture from EHC fire-resistant fluids. These included membrane dry-
ing, vacuum dehydration, and the use of moisture absorbent media.
64 STP 1573 On Fire Resistant Fluids

FIG. 13 Acid number trends after treatment by existing WBA resin A (in wet and dry
form) and two new recommended WBA resins (D&H in dry form).

The active head space membrane dryer utilizes a hollow fiber-based membrane
dryer to generate dry air continuously by separating moisture from a compressed
air source. Whereas, this technology is not suitable for the rapid removal of large
amounts of dissolved water in phosphate ester fluids in a very short duration of
time, it is, however, an excellent option for the cost-effective and continuous re-
moval of moisture from the FRF fluid reservoir and can also be used on an as
required basis (i.e., seasonal use, such as in summer when the relative humidity is
higher than in winter).

FIG. 14 Volume resistivity trends after treatment by existing WBA resins (“A” in wet
and dry form) and two new recommended WBA resins (“D” and “H” in dry
form).
PHILLIPS ET AL., DOI:10.1520/STP157320130102 65

To evaluate this system, four parameters were examined, the relative humidity
of the air entering the dryer (%RH) and the moisture content in the FRF fluid. For
each parameter a high and low value was evaluated. A schematic of the experimen-
tal set up used to evaluate the active head space membrane dryer technology is
shown in Fig. 15.
The used fluid tested had a moisture content of approximately 600 ppm. Addi-
tional moisture was added to the fluid to achieve a moisture content of approxi-
mately 5500 ppm. The results of these tests are summarized in Table 7.
As expected, the use of air with a lower relative humidity was more effective in
removing moisture.
To evaluate the vacuum dehydration technology, fluid was circulated around a
commercial vacuum dehydrator and filtration unit (as shown in Fig. 16).
Approximately 205 l of used fluid was circulated continuously through a com-
mercial vacuum dehydration and filtration unit. Flow rate was 22.7 l/min and the
fluid temperature was kept at 40 C. The circulation was first carried out using a
coarse (>5 lm) filter. This system is very efficient at removing moisture (Table 8).
Other tests were carried out with various zeolite-based molecular sieves and
superabsorbent polymer-based beads to assess moisture removal efficacy from FRF
fluids. A 1:100 ratio of adsorbent to fluid was used in these tests. The media tested
included:
• Molecular Sieves: 3 A, 4 A, and 9 A. These molecular sieves had pore size of 3,
4, and 9 Å respectively.
• Super absorbent beads: SA1 and SA2. SA1 is a polyacrylate-based super ab-
sorbent polymer (SAP) and SA2 is a blend of organo clay and a polyacrylate-
based SAP.
The solution containing the FRF fluid and the treatment product were continu-
ously stirred for a period of 1 and 60 min, respectively. The treatment media was
then separated from the FRF fluid via filtration through a 65 mesh screen. The fil-
trate was then analyzed for moisture, resistivity, acid number, and particle count.
Results of the shakeout tests with a higher contact/stirring time of 60 min are shown
in Table 9. Lower or similar results were obtained for lower contact time of 1 min.
Utilization of molecular sieves and SAPs for treatment of the FRF fluid resulted
in limited improvement in moisture (5–17 %) as well as acid number. However, the
disadvantage of using these media is the increased particle count and reduction in
volume resistivity.
In summary, a vacuum dehydration unit or active head space membrane dryer
are the best options for removing moisture from EHC fluids.

PARTICLE FILTRATION
As indicated earlier, there is significant contamination in the EHC fluids used at
two different CANDU stations and the nature of these particles is the subject of an
ongoing investigation. It appears that the existing ISO Cleanliness Code (which, for
the majority of in-service EHC fluids is at or below 17/15/12) does not reflect the
66
STP 1573 On Fire Resistant Fluids
FIG. 15 Schematic of the membrane dryer (with permission from Kinectrics Inc.).
PHILLIPS ET AL., DOI:10.1520/STP157320130102 67

TABLE 7 Results from the membrane dryer tests.

Initial Moisture Relative Moisture Content Moisture


Content in FRF Humidity, in Reservoir removal
Test Fluid (ppm) RH of air (%) After 1 Week (ppm) (%)

1 646 0 218 66.3


2 612 33.0 6 0.4 586 4.2
3 5550 0 1000 82.0
4 5550 33.0 6 0.4 1420 74.4

presence of very small particles. To evaluate the concentration of particles below


4 lm (the smallest particles measured by the ISO standard), two different methods
were used.
The first method filtered the fluid samples through a 5 lm filter, followed by a
0.45 lm nylon filter. After each filtration the patches were dried before weighing.
The results are presented in Table 10.
The second method utilized an automatic particle counter in which particles
moving through a flow-cell are backlit with a high-output light. The particle images
are collected and then analyzed by computer. The method is claimed to be able to
analyze and size particles as small as 0.7 lm. The results are presented in Fig. 17.
In general, both methods confirmed that the largest concentration of small par-
ticles (particularly in the range of 1 to 4 lm) is at the station with servo valves,
which is also experiencing a problem with micro-dieseling.

FIG. 16 The vacuum dehydration test rig.


68 STP 1573 On Fire Resistant Fluids

TABLE 8 Results from the vacuum dehydration and filtration treatment.

Water Content (ppm)

FRF Fluid As-Received Treatment for 18 h Treatment for 12 h


Treated Fluid Using Coarse Filtera Using Fine Filterb

Used fluid 648 152 134


a
Coarse filter (>5 lm, cyclic stress test/CST rating of 12/07/02) having removal efficiency of 99.9 %.
b
Fine filter (>3 lm, cyclic stress test/CST rating of 08/04/01) having removal efficiency of 99.9 %.

A well–known process for removing small particles is electrostatic filtration.


This involves applying a potential across two parallel plate electrodes between
which the fluid is flowing. Electrophoresis (removal of naturally charged particles
by an electrostatic force) and dielectrophoresis (removal of a charge-neutral particle
by gradient voltage field force) takes place and the particles are collected adjacent to
the electrode.
Two different used EHC fluids (from systems with and without servo valves)
were treated via electrostatic filtration. A Kleentek DOC-R3 EOC filtration system
(15 GPH maximum flow rating) was used with a 60-0471 filter for the test. The
results of electrostatic treatment are presented in Table 11. In each case, approxi-
mately 3 l of fluid was circulated through the electrostatic filter at 0.8 l/min and at
ambient temperature of approximately 23 C for about 24 h. The treated fluid was
then measured for particle size distribution. No patch weights were measured.
It is important to note that these tests were done with a small scale unit, for
a limited test duration and under conditions that were not fully optimized. Fur-
ther improvements in both the particle count and reduction in patch weight are
achievable with an optimized system and longer treatment time.
There are certain advantages in performing such extensive solid particle re-
moval. First, removal of the small particles will minimize the possibility of parti-
cle agglomeration and subsequent plugging of filter elements or accumulating on
critical parts of the EHC system. Second, removal of small particles can reduce

TABLE 9 Results from molecular sieve and super absorbent treatments.

Contact Weight Weight Water Moisture Acid


Treatment Time of Fluid of Mole. Content Removal Number Resistivity
Media (min) (g) Sieve (g) (ppm) (%) (mg KOH/g) (GX-cm)

None NA NA NA 648 – 0.31 3.5


3A 60 102.81 1.57 577 11.0 0.22 2.5
4A 60 125.11 1.72 587 9.4 0.23 2.1
9A 60 98.57 1.18 580 10.5 0.22 1.3
SA1 60 97.74 1.20 597 7.9 0.19 0.9
SA2 60 102.03 1.19 579 10.6 0.20 1.4
PHILLIPS ET AL., DOI:10.1520/STP157320130102 69

TABLE 10 Results of filtration tests with coarse and fine filters.

Weight of Patch After Filtration of 50 ml Through:

Fluid Source 5 -lm Filter 0.45 -lm Filter

Fluid from EHC system without 0.0136 0.0815


servo valves
Fluid from EHC system with 0.0909 Too high to measure
servo valves

the fouling effect of covering the external resin surfaces and thus significantly
improve the efficiency of resin treatment and resin life. Finally, removal of the
submicron particles should also improve the fluid color.
Large-scale tests were performed using an optimized treatment process consist-
ing of electrostatic filtration, moisture removal (by active head space dryer) and
ion-exchange resins. This resulted in significant improvement to fluid characteris-
tics, including color as shown in Fig. 18.

FIG. 17 Particle size distribution in the range 0.7 to 25 lm (left bars are from a station
without servo valves, whereas right bars are from a station with servo valves).
70 STP 1573 On Fire Resistant Fluids

TABLE 11 Particle concentrations before and after electrostatic filtration for only 24 h.

Particle Number of ISO Cleanliness


Station/Fluid Count Particles Code

Used fluid from station with servo 5–10 lm 44 770 18/17/13


valves—as received 10–25 lm 14 930
25–50 lm 882
50–100 lm 126
>100 lm 2

Used fluid from station with 5–10 lm 35 910 18/16/12


servo valves—post-electrostatic 10–25 lm 12 220
treatment 25–50 lm 683
for 24 h 50–100 lm 90
>100 lm 1

Used fluid from station without 5–10 lm 18 370 17/15/12


servo valves—as received 10–25 lm 6015
25–50 lm 358
50–100 lm 52
>100 lm 1

Used fluid from station without 5–10 lm 10 950 16/15/11


servo valves—post-electrostatic 10–25 lm 3955
treatment for 24 h 25–50 lm 182
50–100 lm 18
>100 lm 0

Further Work
Although new information was obtained from the presented study, the need for fur-
ther work was also identified. First of all, the discussed fluid degradation mecha-
nisms produce different acids, some of them weak and some strong. Because
different ion-exchange resins have their preference to absorb one or the other acid
type, it is essential to determine the amount of both strong and weak acids before
selecting resin type for purification treatment. The majority of the current fluid
specifications only require monitoring total acid number.
Another area of concern is a large concentration of small particles, particularly
below 4 lm, that arise in systems where high temperature degradation has taken
place. Because of the current limitation in rating industrial hydraulic fluid cleanli-
ness levels, the concentration of solid particles below 4 lm is unknown. Their pres-
ence, however, may significantly influence the efficiency of the existing purification
treatment and overall EHC system reliability. Application of the latest methodology
for determining particle distribution above 0.7 lm would improve our knowledge
of the amount of small particles present and assist in selecting the most appropriate
filtration process. Because of limitations of the optical method for determination of
PHILLIPS ET AL., DOI:10.1520/STP157320130102 71

FIG. 18 Color improvement of EHC fluid after mechanical and electrostatic filtration,
treatment with IX resins and moisture removal.

particle distribution below 0.7 lm, there is a need for applying other methods to
evaluate particles in the range between 0.1 and 0.7 lm.
In addition, it is important to determine if the resin is releasing any contami-
nants into the fluid to avoid a repetition of the problems identified with fuller’s
earth and activated alumina. Current deposit identification suggests that there is a
difference between commercially available resins of the same type and some of
them may release contaminants into the fluid.
Finally, the interval of ion-exchange replacement requires further review. This
will be a function of the stress level imposed by the EHC system design, operating
conditions and maintenance capability. In addition, it is critical to determine the
required resin volume to maintain efficient and long-term operation between the
resin replacement cycles. All these factors will have a significant impact on both the
EHC system reliability and operating cost.

Conclusions
Ion-exchange resins have become a preferred form of treatment for fire-resistant
fluids used in EHC systems. This treatment appears to be more efficient than other
media in removing acid and can remove other fluid contaminants, for example,
metal soaps and carbon particles. There are, however, several issues related to resin
72 STP 1573 On Fire Resistant Fluids

selection (e.g., which type is most appropriate) and associated purification techni-
ques for moisture control and particle removal, which require further attention. To
ensure that the most effective combination of techniques is used, utilities should
attempt to identify the level/type of contaminants occurring in the EHC fluid. This,
in turn, may help to determine the main degradation mechanisms and lead to sys-
tem modifications that could extend fluid life.
Fluid deterioration is usually a manifestation of system design problems, opera-
tion outside the design envelope or inadequate maintenance. Therefore, improving
system reliability must be done with the co-operation and active engagement of all
stakeholders, including engineering, chemistry, operation and maintenance.

ACKNOWLEDGMENTS
The writers express their appreciation to the CANDU Owners Group for their permis-
sion to share the results of the investigation to date.

References

[1] “Life Cycle Management, Planning Sourcebooks—Vol, 9: Main Turbine Electro-Hydraulic


Control,” Report No. 1009072, EPRI, Palo Alto, CA, 2003.

[2] “Introduction to Nuclear Plant Steam Turbine Control Systems,” Report No. 104885,
EPRI, Palo Alto, CA, 1995.

[3] “Electrohydraulic Control (EHC) Fluid Maintenance Guide,” Report No. 1004554, EPRI,
Palo Alto, CA, 2002.

[4] Phillips, W. D., “The Electrochemical Erosion of Servo-Valves by Phosphate Ester Fire-
Resistant Fluids,” Lubr. Eng., Vol. 44, No. 9, 1988, pp. 758–767.

[5] Phillips, W. D., “The High Temperature Degradation of Hydraulic Oils and Fluids,” J.
Synth. Lubr., Vol. 23, No. 1, 2006, pp. 39–69.

[6] “Steam Turbine Hydraulic Control System, Maintenance Guide,” Report No. 107069,
EPRI, Palo Alto, CA, 1996.

[7] Staniewski, J. W. G., “The Influence of Mechanical Design of Electro Hydraulic Steam Tur-
bine Control Systems on Fire Resistant Fluid Condition,” Lubr. Eng., Vol. 52, No. 3, 1997,
pp. 255–258.

[8] “Electrohydraulic Control Fluid and Elastomer Compatibility Guide,” Report No. 1011823,
EPRI, Palo Alto, CA, 2005.

[9] Wolfe, G. F. and Cohen, M., “Fire Resistant Fluids for Steam Turbine Electrohydraulic
Control Applications: Turbine Lubrication Problems,” ASTM STP, Vol. 437, 1968, pp.
51–72.

[10] Quednau, H., “Entwicklung Nichtbrennbarer Turbinenöle,” Der Erfahrungsaustausch der


Maschinentechnischen Ausschüsse der Vereinigung der Elektrizitätswerke, Meeting
Notes, June 21st, 1944, p. 359.
PHILLIPS ET AL., DOI:10.1520/STP157320130102 73

[11] Phillips, W. D., “Phosphate Ester Hydraulic Fluids,” Handbook of Hydraulic Fluid Technol-
ogy, 2nd ed., G. E. Totten and V. De Negri, Eds., CRC, Boca Raton, FL, 2012, pp. 833–896.

[12] ASTM D2619-09: Standard Test Method for Hydrolytic Stability of Hydraulic Fluids (Bev-
erage Bottle Method), Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA, 2012.

[13] Phillips, W. D. and Sutton, D. I., “Improved Maintenance and Life extension of Phosphate
Esters Using Ion-Exchange Treatment,” J. Synth. Lubr., Vol. 13, No. 3, 1996, pp. 225–261.

[14] Lohrentz, H. J., “Die Entwicklung extrem hoher Temperaturen in Hydrauliksystemen und
die Einflüsse Dieser Temperaturen auf die Bauteile und ihre Funktionen,” Mineralöltech-
nik, Vol. 13, 1968, pp. 14/15.

[15] Anzenburger, J. F., Sr., “Evaluation of Phosphate Ester Fluids to Determine Stability and
Suitability for Continued Service in Gas Turbines,” Lubrication Engineering, Vol. 43, No.
7, 1986, pp. 528–532.

[16] FMC Corporation, U.K. (unpublished).

[17] Roberton, R. S. and Allen, J. M., “A Study of Oil Performance in Numerically Controlled
Hydraulic Systems,” Nat. Conf. Fluid Power, Vol. 28, 1974, pp. 435–454.

[18] Lhomme, V., Bruneau, C., Soyer, N., and Brault, A., “Thermal Behavior of Some Organic
Phosphates,” Ind. Eng. Chem. Prod. Res. Dev., Vol. 23, No. 1, 1984, pp. 98–102.

[19] Schober, J., “Fire-Resistant Hydraulic Fluids,” Brown Boveri Rev., Vol. 53, No. 12, 1966,
pp. 142–147.

[20] Tersiguel-Alcover, C., “Traitement et Contrôle des Fluids de régulation Difficilement


Inflammables (Type Ester Phosphate) en Fonctionnement,” Report No. P33/4/4200/78-
42, EDF, New York, 1978.

[21] Tersiguel-Alcover, C., “La Filtration des Esters Phosphates sur Alumine Active,” Report
No. P33/4200/81-24, EDF, New York, 1981.

[22] Grupp, H., “Aufbau von schwer entflammbaren Hydraulikflüssigkeiten auf Phosphorsäur-
eesterbasis, Erfahrungen aus dem praktischen Einsatz im Kraftwerk [Construction of
fire-resistant Hydraulikflussigkeiten on phosphoric acid ester-based experiences from
practical use in power station],” Der Masch. Schaden, Vol. 52, No. 3, 1979, pp. 73–77 (in
German).

[23] Grupp, H., “Betriebsbedingte Veränderungen von Schwer Entflammbaren Hydraulikflüs-


sigkeiten und Deren Analytischer Nachweis [Changes in fire-resistant fluids induced by
operation and their analytical detection],” VGB Kraftwerkstechnik., Vol. 72, No. 7, 1992,
pp. 630–635 (in German).

[24] Phillips, W. D., “The Conditioning of Phosphate Esters in Turbine Applications,” Lubr.
Eng., Vol. 39, No. 12, 1979, pp. 766–778.

[25] Habermas, R. E., Howard, L. R., and LaVere, D. J., “In-Service Reconditioning of Turbine
Generator Hydraulic Fluid,” Power Engineering, Vol. 28, PennWell Publishing Company,
Fair Lawn, NJ, 1982.

[26] “Composite Adsorbent for Removing Acids From Organophosphate Functional Fluids,”
U.S. Patent No. 4,751,211 (1986).
74 STP 1573 On Fire Resistant Fluids

[27] Wolfe, G. E. and Whitehead, A., “Experience With Phosphate Ester Fluids as Industrial
Steam Turbine Generator Lubricants,” Lubr. Eng., 1977, Vol. 34, No. 8, pp. 413–420.

[28] Brandt, F. C. J., Braun, D., and Trost, R., “Erste Erfahrungen mit Einem neuen Reinigungs-
verfahren für Synthetische Schwerbrennbaren Hydraulikflüssigkeiten [First experiences
with a new cleaning method for synthetic heavy combustibles hydraulikflussigkeiten],”
Der Masch. Schaden, Vol. 57, No. 5, 1984, pp. 194–196 (in German).

[29] Braun, D., “Aufbereitung von Phosphatester mit Ionenaustauscherharzen: Historischer


Überblick,” Symposium on Advanced Maintenance Practices for Modern Power Genera-
tion Applications, Hydac GmbH, St. Ingbert, Germany, Jan 21–22, 2009.

[30] Sand, H., “Regeneration schwerbrennbarer Flüssigkeiten [Regeneration of difficult flam-


mable liquids],” Technische Akademie Esslingen Course 20167/68.380, Synthetische
Flüssigkeiten als Hydraulik und Schmieröle in Turbinenlagen, Expert Verlag, Renningen,
Germany, 1988 (in German).

[31] Phillips, W. D. and Sutton, D. I., “Improved Maintenance and Life Extension of Phosphate
Esters Using Ion-Exchange Treatment,” J. Synth. Lubr., Vol. 13, No. 3, 1996, pp. 225–261.

[32] Dufresne, P. E., “On-Line Remediation of Ester Based Synthetic Oils,” Lubr. Eng., Vol. 6,
2001, pp. 16–20.

[33] Austin, E. M., Gauthier, C., and Staniewski, J. W. G., “Operating Experience With Ion
Exchange Purification Systems for Steam Turbine Control Fluids,” Society of Tribologists
and Lubricating Engineers Annual Meeting, Cincinnati, OH, May 20–23, 1996, pp.
211–224.

[34] Staniewski, J. W. G., “Maintenance Practices for Steam Turbine Control Fire Resistant
Fluids, Part 1,” J. Synth. Lubr., Vol. 23, No. 3, 2006, pp. 109–121.

[35] Staniewski, J. W. G., “Maintenance Practices for Steam Turbine Control Fire Resistant
Fluids, Part 2,” J. Synth. Lubr., Vol. 23, No. 3, 2006, pp. 123–135.

[36] Collins, K. G. and Duchowski, J. K., “Effectiveness of the Ion Exchange/Vacuum Dehydra-
tion Treatment of Phosphate Ester Fluids,” J. Synth. Lubr., Vol. 19, No. 1, 2002, pp. 31–38.

[37] “Novel Options for Improved Maintenance and Life Extension of Steam Turbine Control
Fire Resistant Fluids,” Technical Note 11-4009, CANDU Owners Group (COG), Toronto,
Canada, 2011.

[38] “Evaluation of Various Technologies for Improved Treatment of Phosphate Ester Fluids,”
Technical Note-12-4006, CANDU Owners Group (COG), Toronto, Canada, 2013.
FIRE RESISTANT FLUIDS 75

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130106

Matthew G. Hobbs1 and Peter T. Dufresne Jr.1

Phosphate Ester-based Fluid


Specific Resistance: Effects of
Outside Contamination and
Improvement Using Novel Media
Reference
Hobbs, Matthew G. and Dufresne, Peter T. Jr., “Phosphate Ester-based Fluid Specific
Resistance: Effects of Outside Contamination and Improvement Using Novel Media,” Fire
Resistant Fluids, STP 1573, John Sherman, Ed., pp. 75–92, doi:10.1520/STP157320130106,
ASTM International, West Conshohocken, PA 2014.2

ABSTRACT
The ASTM D1169-11 test method is currently used to determine the continued
functionality of operational, fire-resistant phosphate ester (PE) fluids. The theory
behind this is that greater resistance will reduce the propensity of lubricating
fluids to cause streaming-current corrosion; however, to the average PE fluid
user, the method provides a quantitative value that describes the overall general
cleanliness of PE fluids. Specific PE contamination, such as by acids and water,
negatively affects (lowers) the bulk volume specific resistance. The specific
resistance can potentially fall below the condemning limit set by the PE fluid
provider(s) and the original equipment manufacturer (5 GX  cm) if such
contamination levels are high. Conversely, there is contamination that will
increase the resistance of operational PE fluids; this contamination is in the form
of extremely small, abrasive, dielectric aluminosilicate particles. This
contamination comes from the manufacturer-recommended PE purification
media Selexsorb GT. The dielectric nature of the particulate gives the user, and
ASTM D1169-11, the appearance of resistance improvement, but particle
abrasiveness is concomitantly catalyzing fluid degradation and causing undue
mechanical wear. Data documenting the misleading PE fluid resistance
improvement induced by aluminosilicate contamination are described. Novel

Manuscript received June 10, 2013; accepted for publication November 19, 2013; published online April 4,
2014.
1
EPT, Calgary, AB, Canada T2B 3R4.
2
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
76 STP 1573 On Fire Resistant Fluids

methods to improve specific resistance without any deleterious side effects also
are presented to help PE fluid users avoid condemning entire reservoirs as a
result of low specific resistance.

Keywords
resistivity, contamination, fluid treatment

Introduction
Phosphate ester–based hydraulic fluids are synthetic, nonaqueous liquids, generally
comprising triaryl phosphate esters. The unique properties of these fluids, including
their superior fire resistance and ability to self-extinguish, have made them the fluid
of choice for applications in which the fluid is likely to come into contact with
extremely high temperatures or pressures [1]. These conditions are met in the elec-
trohydraulic control (EHC) systems involved in power plant operation, and as a
result, phosphate esters are the hydraulic fluid most commonly used in this
application.
The specific resistance, or resistivity, of a phosphate ester–based fluid is a mea-
sure of its electrical insulating properties and is inversely proportional to the fluid’s
ability to conduct electrical charge. Phosphate esters are inherently electrically insu-
lating, and any conductivity observed generally arises as the result of ionic or ion-
forming contaminants. Examples of common ionic or potentially ion-forming con-
taminants include acids, dissolved metals, chlorides, and water. The presence of
these contaminants deleteriously affects (lowers) the resistivity of the fluid, making
the production of streaming currents possible when the fluid flows. Such streaming
currents are known to produce damaging corrosion in EHC systems [2]. In highly
resistive (very poorly conductive) fluids, the streaming current magnitude is negligi-
ble and the associated corrosion is decreased or eliminated. Fluid manufacturers
and original equipment manufacturers (OEMs) therefore recommend maintaining
resistivity above a value of 5 GX  cm. The resistivity of phosphate ester–based flu-
ids is monitored according to ASTM D1169-11 [3]. In addition to providing an in-
dication of the potential for streaming current corrosion, this value is frequently
interpreted as a general measure of the fluid’s overall cleanliness, with high resistiv-
ity values being associated with fluids in good condition.
In an effort to maintain the phosphate ester’s condition and ensure maximum
equipment reliability, fluid manufacturers recommend the use of an off-line acid and
particulate filtration system; the most common of these systems involve the use of
Fullers Earth or Selexsorb GT as adsorbent acid-scavenging media.3 These are gener-
ally thought to remove the contaminants that deleteriously affect phosphate ester re-
sistivity; however, they can also release other contaminants into the fluid. The

3
Selexsorb is a registered trademark of BASF.
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 77

propensity of Fullers Earth to release resistivity-decreasing dissolved metals (most fre-


quently calcium, magnesium, and iron) that have the potential to form metal soaps
that lead to equipment-damaging solid deposits and sludges is well established [4,5].
This study therefore focused on the effects of Selexsorb GT contamination on phos-
phate ester (PE) fluid conditions. The use of other purification media, specifically, ion
exchange resins, has been attacked as ineffective at maintaining fluid resistivity above
the condemning limit [6]; however, data presented herein demonstrate that ion
exchange can be used to significantly improve the condition of a PE fluid.

The Effects of Selexsorb GT and Ion


Exchange–based Purification
Media upon the Resistivity of Phosphate
Ester Fluids
In order to compare the efficacy of Selexsorb GT and ion exchange EHC fluid ad-
sorbent purification media, we set out to perform laboratory-scale treatments of a
commercially available PE-based fire-resistant fluid. Our initial experimental setup
involved passing the unused fluid through different filter media contained within a
standard glass chromatography column. The adsorbent media were supported by a
built-in coarse fritted filter at the base of the column. Following treatment, the resis-
tivity of all fluids was determined at 20.0 C 6 0.3 C using a Quadtech 1865 MX
meter and a temperature-controlled bath according to the procedures outlined in
ASTM D1169-11 [3].
When the fluid (300 ml) was passed through Selexsorb GT (four reservoir turn-
overs), its resistivity improved by more than a factor of 10, increasing from a value
of 6.20 GX  cm in the untreated fluid to a value of 71.72 GX  cm. These initial data
suggest that Selexsorb GT is a highly efficient filtration medium for the improve-
ment of EHC fluid resistivity. Although the resistivity increase of the same fluid
treated by a proprietary ion exchange medium was impressive (485 % increase), it
was less than half of that observed with Selexsorb GT (1057 %).
Wishing to further explore the effects of Selexsorb GT treatment upon PEs, we
devised an improved experimental setup capable of demonstrating the effects of
increased reservoir turnovers using a machined steel column and a pump to force
the fluid through the purification media at a rate of 20 ml/min. A small amount of
cotton wool was placed at the base of the column to support the adsorbent media.
Interestingly, after 48 h (approximately 115 reservoir turnovers), a significantly
smaller increase in fluid resistivity was observed (477 % increase) relative to the
1057 % increase observed earlier after only four reservoir turnovers.
Initially this was puzzling, as we had anticipated enhanced resistivity improve-
ment with a greater number of reservoir turnovers. We therefore performed addi-
tional standard PE analyses in order to determine the root cause of this difference.
Through the course of these analyses, we determined that the most significant
78 STP 1573 On Fire Resistant Fluids

difference between the two Selexsorb GT–treated fluid samples was their particle
count. Whereas the particle count of the fluid that was passed through the pressur-
ized column increased only slightly, from an initial count of 18/16/13 to a value of
21/17/11 following treatment (115 turnovers), the particle count of the fluid passed
through the standard chromatography column increased significantly to a post-
treatment (four turnovers) value of 24/22/17. Following examination of the distinct
column setups used, we determined that the cause of this difference in particle
count was the adsorbent media support used in each. In the standard chromatogra-
phy column used, the fritted disc filter was relatively coarse (40 -lm to 60 -lm pore
size) and would therefore be expected to allow a significant number of fines to pass
into the fluid [7]. The cotton wool used in the pressurized column offered a finer
degree of filtration [8].
The above results suggest that, in the absence of a fine mechanical filter, Selex-
sorb GT introduces particulate contamination into EHC fluids. Indeed, even during
the experimental setup when the white crystalline Selexsorb GT was poured care-
fully into both columns, a fine cloud of white dust remained suspended in the air
for some time; undoubtedly, this fine sediment leaches into fluids passing through
the adsorbent. In order to assess the size and abrasive potential of Selexsorb GT, we
examined it using a microscope at 4 and 120 magnification (Fig. 1). The images
in Fig. 1 demonstrate the irregular and sharp shape of Selexsorb GT crystals. Note
also that the image magnified 120 displays not only one large grainy particle but
many smaller such particles that are barely visible even under 120 magnification.
Selexsorb GT is a patented mixture of alumina and Y-zeolite, but our own x-ray
diffraction analysis suggests that it is a zeolitic aluminosilicate mineral
(Na2Al2Si4O12  8H2O). Aluminosilicate minerals are abrasive and have Mohs hard-
ness values greater than that of steel, and thus their presence in an EHC system
would be expected to result in significant wear [9].

FIG. 1 Selexsorb GT grains as seen under 4 optical magnification (left) and 120
scanning electron microscopy magnification (right). Note that even under 120
magnification, there are still extremely fine particles covering the surface of the
grain. Some of these particles have broken away from the grain and formed
extremely fine solids (visible in the upper right of the photo).
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 79

THE EFFECT OF SOLID SELEXSORB GT–RELATED CONTAMINATION UPON


PHOSPHATE ESTER RESISTIVITY
Zeolite materials such as Selexsorb GT are well-known dielectric species that resist
current flow [10]. We were therefore curious to determine the effect that fine, sus-
pended Selexsorb GT contamination would have upon the resistivity of PE-based
fluids (Fig. 2) and whether this contamination could account for the unexpected
differences in resistivity observed between the Selexsorb GT–treated fluids filtered
through a coarse fritted filter and those filtered through a finer cotton wool plug.
In order to assess the effect of this fine aluminosilicate contamination, we per-
formed an additional filtration upon the EHC fluid that had been treated by pass-
ing it through the Selexsorb GT supported by a coarse fritted filter. Recall that the
original Selexsorb GT treatment resulted in a 1057 % increase in the resistivity of
the fluid in addition to a significant increase in the fluid’s particle count. After
this Selexsorb GT–treated fluid had passed through a 1.2 -lm patch, the resistivity
of the treated fluid was found to have decreased by 76 % to a final value of
17.46 GX  cm. In an effort to confirm that the fine particulate contamination
removed during this additional filtration was responsible for the initially observed
resistivity enhancement and that the method used for the secondary filtration did
not somehow deleteriously affect the fluid’s resistivity, we repeated the additional
filtration using the PE fluid that had been purified by means of ion exchange.
Unlike the Selexsorb GT–treated fluid, following mechanical filtration, the resis-
tivity of the fluid treated via proprietary ion exchange resin increased slightly
(7 %). These observations demonstrate that the resistivity improvement that
occurs during Selexsorb GT filtration of PE fluids is related to the introduction of
potentially abrasive solid contamination and that the resistivity improvement that
arises as a result of this particulate is reversible. These findings support our origi-
nal hypothesis that the difference in the filtering abilities of the coarse fritted disc

FIG. 2 Effect of phosphate ester fluid treatment on specific resistance.


80 STP 1573 On Fire Resistant Fluids

and the cotton wool plug employed was responsible for the differences in fluid re-
sistivity noted earlier.
Although the above data suggest that the presence of fine particulate contamination
released by Selexsorb GT into the fluid is responsible for a significant, but reversible,
increase in its measured resistivity, there are many factors that influence bulk resistivity
in PEs, including acidity, dissolved metals, chloride content, and moisture content. In
an effort to unequivocally demonstrate that the observed resistivity enhancement arose
either solely or primarily as the result of the abrasive particulate contamination intro-
duced by Selexsorb GT, we performed additional analyses for each of the above-
discussed factors on a virgin fluid sample and on the fluid following four passes through
a glass chromatography column filled with Selexsorb GT according to the following
methods: total acid number according to ASTM D664 [11], dissolved-metals analysis
according to ASTM D5185 [12], chloride content according to a modified version of
ASTM D4929 [13], and moisture according to ASTM D7546 [14]. (Note: because of the
high boiling range of PEs, no distillation to obtain the naphtha cut was performed. Test-
ing was conducted on the entire sample. Because the PE tested did not contain hydrogen
sulfide, no caustic washing step was performed. The inorganic chlorides present in the
water washings of the sample were analyzed separately from the organic chlorides,
which remained in the organic phase. The result reported is the sum total chloride [inor-
ganic and organic] concentration present in the original sample.) We hoped that by
assessing the results obtained (Table 1) we could conclusively identify the mechanism by
which Selexsorb GT treatment improves the resistivity of PEs.

ACID NUMBER AND ITS INFLUENCE UPON SPECIFIC RESISTANCE PRIOR


TO AND FOLLOWING SELEXSORB GT TREATMENT
As anticipated, the acid number of the fluid was halved following treatment with
the acid scavenger Selexsorb GT, decreasing from 0.08 mg KOH/g in the virgin
fluid sample to 0.04 mg KOH/g. Because acids (particularly strong acids) are ioniz-
able species in solution, this decrease would result in a concomitant increase in the
resistivity of the fluid (as observed). The magnitude of the observed resistivity
increase, however, was much greater than would be expected based upon the
observed acid number decrease. Treatment under identical conditions with a

TABLE 1 Effect of phosphate ester treatment on fluid resistivity and other properties upon which
resistivity is dependent.

PE Fluid Treatment None Selexsorb GT Ion Exchange

Resistivity, GX  cm 6.20 71.72 36.28


ISO particle count 18/16/13 24/22/17 20/18/13
Total acid number (TAN), mg KOH/g 0.08 0.04 0.04
Dissolved metals, ppm 1.4 7.3 0.7
Chlorides, ppm 141.5 130.6 119.2
Moisture, ppm 667 1301 693
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 81

proprietary ion exchange resin, for instance, resulted in the same acid number
reduction; however, the resistivity of the treated fluid increased by 485 % (com-
pared to the 1057 % increase observed following Selexsorb GT treatment). In an
effort to demonstrate that the presence of fine particulate contamination was re-
sponsible for an exaggeration of the resistivity of Selexsorb GT–treated fluids and
that this improvement did not occur solely due to the medium’s acid-scavenging
ability, we performed acid number testing on the fluid following Selexsorb GT
treatment and then again following the filtration used to remove the particulate
contamination introduced. Although the filtration decreased the resistivity of the
PE by removing fine dielectric contaminants, it had no effect upon its acid number.

METAL CONTENT AND ITS INFLUENCE UPON SPECIFIC RESISTANCE PRIOR


TO AND FOLLOWING SELEXSORB GT TREATMENT
The presence of dissolved metals, which exist in their ionic forms in solution, is
expected to deleteriously affect (decrease) the bulk resistivity of EHC fluids. We
were therefore curious about whether Selexsorb GT demonstrated any ability to
remove metals from solution. Metals analysis of the virgin and treated fluids, how-
ever, demonstrated that Selexsorb GT treatment increased the concentration of total
dissolved metals from 0.8 ppm in the virgin sample to 7.1 ppm in the treated sample.
The majority of this increase was attributed to the observed increase in aluminum
levels, from 0 ppm prior to treatment to 7.0 ppm following treatment. This finding
confirms that Selexsorb GT releases potentially abrasive aluminum-containing con-
taminants into solution during its use. This increase in dissolved metals cannot
explain the observed increase in fluid resistivity; indeed, such an increase in metal
content would be expected to reduce, not increase, the specific resistance.
It is interesting to note that the increase in dissolved metals following Selexsorb
GT treatment discussed above occurred under extremely mild conditions (four reser-
voir turnovers at room temperature under 1 atmosphere pressure). When subjected
to more aggressive conditions (sonication), Selexsorb GT was found to significantly
contaminate PE fluid. Following sonication for 8 h at a temperature of 35 C, the total
dissolved metals present in the PE fluid increased by 522.4 ppm. The vast majority of
this increase was the result of increases in the dissolved aluminum (426.5 ppm) and
silicon (95.0 ppm) contents of the fluid. For the purpose of comparison, a fluid
treated by proprietary ion exchange resin under identical conditions did not increase
the concentration of metals present. In fact, sonication in the presence of the ion
exchange medium was actually found to slightly decrease the level of total dissolved
metals present. This result demonstrates the concerning potential for abrasive alumi-
nosilicate contamination that exists when Selexsorb GT is used in an attempt to
purify PE-based fluids. In addition to the potential for abrasive wear, the presence of
metals such as those noted above is known to significantly increase the rate of PE
degradation through catalysis [15]. Consequently, the presence of soluble and insolu-
ble metals released into solution by Selexsorb GT is expected to accelerate equipment
wear while simultaneously damaging the condition of the fluid.
82 STP 1573 On Fire Resistant Fluids

CHLORIDE CONTENT AND ITS INFLUENCE UPON SPECIFIC RESISTANCE


PRIOR TO AND FOLLOWING SELEXSORB GT TREATMENT
The relationship between the presence of chlorides and decreased PE specific resist-
ance is well established [16]; therefore, the chloride content of the fluid (both organic
and inorganic) was analyzed prior to and following Selexsorb GT treatment. Selexsorb
GT treatment was found to remove a small percentage of the chloride present in the
PE fluid (8 %); however, identical treatment of the same fluid using an ion exchange
medium proved to be twice as efficient, removing 16 % of the chlorides present in the
fluid. Although the observed decrease in chloride levels can account for an increase in
the resistivity of the treated PE, it is insufficient to account for the magnitude of the
specific resistance increase observed. Indeed, based upon the changes observed in
fluid chloride content, one would expect to see a greater increase in resistivity
following ion exchange treatment. Clearly, the exaggerated resistivity improvement
provided by Selexsorb GT cannot be accounted for on the basis of chloride removal.

MOISTURE CONTENT AND ITS INFLUENCE UPON SPECIFIC RESISTANCE


PRIOR TO AND FOLLOWING SELEXSORB GT TREATMENT
Finally, the moisture content of the virgin EHC fluid and that of the same fluid follow-
ing Selexsorb GT filtration was determined. The water content of the as-received virgin
EHC fluid was found to be 667 ppm, and following treatment of the fluid with Selex-
sorb GT, this value was found to have increased by 95 % to 1301 ppm. Although water
itself is a dielectric material, its polar nature makes it an excellent solvent for ionic spe-
cies; the presence of these species decreases the resistivity of PEs, meaning that changes
in the water content of the fluid are not responsible for the observed resistivity increase
following treatment with Selexsorb GT. The ingression of water into the EHC fluid
during the course of its exposure to Selexsorb GT is likely a consequence of the chemi-
cal composition of this purification medium, as aluminosilicate minerals (Na2Al2-
Si4O128H2O) incorporate water into their crystalline structures. Although these
species are hygroscopic in their dehydrated state, once hydrated they may release water
into the fluid. Treatment of the same EHC fluid, under identical conditions, with a
proprietary ion exchange resin resulted in only a modest increase (4 %) in the moisture
level of the fluid. This is interesting given that the use of ion exchange resins has been
attacked because of their tendency to contaminate PEs with water [4]. This misconcep-
tion is due to the fact that ion exchange resins, as received from the manufacturer, are
wet and contain up to 55 % water. Although it is possible for water ingression to occur
if the ion exchange resin is used as received, ion exchange resins can be readily dried in
a controlled and reproducible manner. The very small amount of water added to the
EHC fluid during the ion exchange experiment conducted above indicates that the pro-
prietary resin employed was likely pretreated in this manner.

THE MECHANISM OF PHOSPHATE ESTER RESISTIVITY IMPROVEMENT


THROUGH SELEXSORB GT TREATMENT
The additional analyses conducted above (acid number, dissolved metals, chlorides,
and moisture content) unequivocally demonstrated that the enhanced resistivity
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 83

improvement observed when treating PE-based fluids with Selexsorb GT was due
to the presence of fine dielectric aluminosilicate contamination. Of the changes to
the above parameters that are associated with resistivity improvement, only those
observed for the acid number of the fluid filtered through Selexsorb GT can
account for any increase in its resistivity. This improvement in acid number was
subsequently shown to be insufficient to account for the exaggerated resistivity
observed.
Conventional thinking holds that specific resistance is influenced by the pres-
ence of soluble contaminants in PE fluids (acids, dissolved metals, chlorides, and
water); however, the results presented herein demonstrate that insoluble species can
also exert a significant influence. The effect of insoluble dielectrics upon the resistiv-
ity of geological fluids is summarized by Archie’s law (Eq 1) and is well established
[17]. According to this law, the resistivity of the dielectric solid and fluid-
containing system (RO) is proportional to the fluid’s resistivity (RF). In geology, the
constant of proportionality is described as the formation factor (F). This factor
describes the degree to which the fluid is diluted by electrically insulating mineral
grains, and its value is always greater than 1. When there is no dielectric solid pres-
ent in a system, the formation factor is equal to unity and the specific resistance of
the system (RO) is equal to that of the fluid (RF). As the amount of dielectric miner-
als present in the system increases, the path taken by currents in the fluid becomes
less direct and current flow is, to a degree, impeded. As a result, the magnitude of
the factor F increases and the resistivity of the system (RO) is therefore also
increased (even when there is no change to that of the fluid, RF).

RO ¼ FRF : (1)

where:
RO ¼ resistivity of a system containing a fluid and an insoluble dielectric
medium,
F ¼ formation factor describing the degree of insoluble species present in the
system, and
RF ¼ resistivity of the fluid in the system of interest.
In geology, the fluid is water and the system is water in a porous mineral me-
dium. When this law is applied to EHC systems, it is clear that the “fluid” involved
is the EHC fluid and all of the soluble contaminants contained therein, and the
“system” is the combination of the EHC fluid, soluble contaminants, and insoluble
contaminants. Because the value of the term F in Eq 1 is always greater than 1 in
the presence of insoluble dielectric contaminants, the aluminosilicate minerals
released into PE fluids during Selexsorb GT treatment will increase the resistivity of
the system.
Because the resistivity enhancement that occurs as a result of solid Selexsorb
GT contamination was demonstrated to be reversible when the solid contami-
nants were removed via filtration, the exaggerated resistivity measurements
84 STP 1573 On Fire Resistant Fluids

obtained following EHC fluid treatment with Selexsorb GT should not be


regarded as meaningful. In real-world applications, Selexsorb GT filters are typi-
cally used in conjunction with mechanical filters; however, the mechanical filters
employed are not as efficient as the nylon filter patch (1.2 lm) employed in our
laboratory study. The mechanical filters suggested for use in this application have
a b ratio of b3 ¼ 200, meaning that they are 99.5 % efficient in the removal of
particles 3 lm in size [4]. In an effort to demonstrate that fine Selexsorb GT–-
derived contamination remained even following filtration through a 1.2 -lm filter
patch, we performed an additional patch weight analysis using a finer 0.45 -lm
nylon patch. Indeed, the patch weight obtained for the finer 0.45 -lm filtration
(7.2 mg/50 ml) was nearly identical to that obtained during the coarser 1.2 -lm fil-
tration (7.9 mg/50 ml), suggesting that there were as many fine particles between
0.45 lm and 1.2 lm in size as there had been particles larger than 1.2 lm in the
original Selexsorb GT–treated solution. Undoubtedly, ultrafine particles smaller
than 0.45 lm remained in solution even following both filtrations. Although the
mechanical filters in place in real-world EHC systems might be sufficient to
remove the larger, coarser particles introduced by Selexsorb GT, they are insuffi-
cient to remove the fine and ultrafine dielectric and potentially abrasive alumino-
silicates introduced simultaneously.
In all of the studies conducted, ion exchange resins were found to be at least
as effective as, and in many cases more effective than, Selexsorb GT at improving
the condition of PE-based EHC fluids. There are many different classes of ion
exchange resins and, indeed, thousands of different specific resins, so we turned
our attention to the study of these media with the aim of establishing which
classes and which specific resins were best suited for in-service EHC fluid
purification.

Improvement of In-service Phosphate


Ester–based EHC Fluid Specific Resistivity Using
Ion Exchange Adsorbent Purification Media
In an effort to better understand the effects of ion exchange resins upon the
specific resistance of in-service EHC fluids, we collected samples of 50 different
PE fluids from numerous real-world EHC operators. The initial resistivity values
of these fluids were measured in the same manner as those of the fluids dis-
cussed above. The effects of various ion exchange media upon the different flu-
ids’ resistivities were then examined after passing the fluids through laboratory-
scale columns. All together, the 50 different fluids were passed through 70 dif-
ferent columns containing 1 of 11 distinct ion exchange resins. As a result, we
are confident that this project employed a sufficiently large sample size and that
general conclusions can therefore be drawn from the results presented below
(see the Appendix for a full table of results).
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 85

Of the 50 different fluids tested, 84 % showed some improvement in resistiv-


ity following treatment by at least one specific ion exchange resin. Moreover, the
specific resistance of 74 % of the fluids examined improved regardless of the iden-
tity of the resin used. Of the samples taken, 30 were received in poor condition
with initially measured resistivities below the condemning limit of 5 GX  cm sug-
gested by most OEMs. Despite the claim that “it is difficult to keep the resistivity
up above the 5 GX  cm limit (using ion exchange)” [6], fluid resistivity was
indeed restored to a value above this limit in 17 of these 30 heavily contaminated
fluids. Of the 70 column trials completed, improvements in fluid specific resist-
ance were observed in 81 % of cases. The corollary to this observation is that in
19 % of the column trials, no improvement or, worse still, a negative effect on re-
sistivity was observed. The average magnitude of this negative effect, however,
was relatively small; in cases where resistivity decreases were observed following
ion exchange treatment, the average decrease was 1.83 GX  cm. The average resis-
tivity increase (observed in 81 % of the trials performed) was of a significantly
larger magnitude (4.11 GX  cm). In two trials, no change in specific resistance
could be measured because of the very poor condition in which the in-service flu-
ids were received, suggesting that there are limits on the remediating capabilities
of ion exchange resins. We do note, however, that an in-service EHC fluid with a
resistivity as poor as 0.67 GX  cm was effectively treated with ion exchange
media; the resistivity of this fluid following treatment was determined to be
16.64 GX  cm.
Although our study suggests that ion exchange resins can be used to effectively
control the resistivity of EHC fluids, we must note that some specific resins and
some specific types of resins are more effective than others (Fig. 3). Indeed, ion
exchange resins are not all created equally. Because of the proprietary nature of the
resins studied, we are unable to reveal the specific identities of resins (identified
here only by a Greek letter) or resin types (identified only by a letter). As a result,
the discussion below is somewhat limited. We do, however, wish to single out the
exceptional resistivity control provided by proprietary ion exchange resins of types
B and E.

TYPE B ION EXCHANGE RESINS


The results obtained demonstrate that type B resins were a highly effective class
of ion exchange resins for resistivity control. On average, type B resins were able
to increase the resistivity of the test fluids by 4.05 GX  cm. More impressively,
the greatest single resistivity improvement observed during the course of this
study was noted following treatment with a type B resin (c). In this particular
case, treatment of the fluid by c was able to increase the fluid’s specific resistance
by 17.3 GX  cm, an improvement of 118 %. Indeed, three of the five top-
performing resins examined (c, d, and e) were type B resins used without prior
modification.
86 STP 1573 On Fire Resistant Fluids

FIG. 3 In-service phosphate ester fluid specific resistance improvements following


treatment with various ion exchange purification media.

TYPE E ION EXCHANGE RESINS


Although various types of ion exchange resins (type B in particular) were found to
be effective at improving PE resistivity without modification, the most effective ion
exchange purification media tested during the course of this project were the altered
type E resins. Among the in-service fluids treated using this proprietary resin type,
the average resistivity improvement observed was an impressive 7.93 GX  cm. The
magnitude of this improvement is significantly greater than that observed for any
of the other classes of resins studied. Moreover, of the 70 column trials completed
on the 50 different in-service EHC fluids, the second-most significant resistivity
enhancement observed for any one trial occurred with the use of a type E resin (k).
The resistivity of this particular fluid jumped from 7.78 GX  cm upon receipt to
23.92 GX  cm following treatment with the proprietary resin, an improvement of
16.14 GX  cm (207 %).
Overall, the results obtained demonstrate the efficacy of ion exchange media
in the control of real-world in-service PE-based EHC fluid resistivities. More-
over, these fluids demonstrated this ability without the detrimental effects
observed in fluids treated with more conventional adsorbent purification media
such as Fullers Earth and Selexsorb GT. Although it is well known that Selex-
sorb GT is often incompatible with EHC fluids that have been previously puri-
fied using other media (such as Fullers Earth or activated alumina) [6], the
above data demonstrate that ion exchange resins are broadly applicable for PE
resistivity control and, indeed, general condition improvement.

Conclusion
Selexsorb GT initially appeared to be an effective adsorbent purification medium
for PE fluids based upon the dramatic improvements in fluid resistivity that were
observed following treatment. Upon further examination of the data and after addi-
tional testing, however, we found that the resistivity improvements observed
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 87

following Selexsorb GT treatment were misleading. These resistivity


“improvements” are reversible and occur as the result of fine dielectric alumi-
nosilicate contamination introduced into the fluid by the adsorbent medium.
Although, to the best of our knowledge, this phenomenon has not previously
been noted in PEs, it is well established in geological fluids. The suspended
dielectric solids appear to impede current flow through the fluid, resulting in
the appearance of enhanced fluid resistivity. This aluminosilicate contamina-
tion, in addition to producing exaggerated resistivity improvements, deleteri-
ously affects the fluid’s condition, as the metal-containing particles (primarily
aluminum and silicon) released by Selexsorb GT can give rise to accelerated
wear and fluid degradation rates. Whereas aluminosilicate contamination yields
increases in resistivity, other forms of contamination (acids, metals, chlorides,
moisture, etc.) yield decreases. The interplay among these factors (and poten-
tially others) is complex, and the measured resistivity of a PE should therefore
not be viewed as a simple overall measurement of fluid cleanliness. Although
the measurement of specific resistance in PE fluids remains important as a
means of assessing the potential for damaging streaming current corrosion,
many other analyses should also be performed in order to accurately assess
the condition of the fluid. As with other fluid condition monitoring and oil
analysis programs, end users can make the best decisions to increase the reli-
ability and performance of their systems by carefully weighing and interpreting
the results obtained from many different test methods. Selexsorb GT users
should exercise caution when interpreting the results of resistivity testing per-
formed upon their fluids.
Although ion exchange media are a well-known alternative to Selexsorb GT pu-
rification in PE applications, their propensity to release water into the fluid and
their ability to control PE specific resistance have been criticized. With regard to
the former of these two criticisms, the moisture contents of fluids being treated via
ion exchange can be readily managed by taking appropriate measures to ensure that
the resins have been sufficiently dried prior to use and that water ingression does
not occur during service. The results presented herein demonstrate that the latter of
these criticisms (that ion exchange resins are ineffective at controlling fluid resistiv-
ity) is erroneous. When compared to Selexsorb GT, treatment of PE-based fluids
using ion exchange media proved to be equally effective at reducing the fluid’s acid
number and more effective at producing meaningful improvements in fluid resistiv-
ity. Moreover, treatment via ion exchange did not result in the introduction of
potentially fluid-damaging metals or wear-producing fine solids. A comprehensive
study involving 11 different ion exchangers, 50 different in-service PE fluids, and 70
separate trials demonstrated the effectiveness of ion exchange as a technique to
improve EHC fluid resistivity in real-world applications. The results of this study
suggest that ion exchange resins are not all created equal. Indeed, the best
performance was observed during trials employing proprietary ion exchange resins
of type E.
88
Appendix

STP 1573 On Fire Resistant Fluids


TABLE 2 Effect of ion exchange treatment using different resins (types A through E) on in-service PE resistivity (GX  cm).

Fluid Fluid Type Initial a (D) b (D) c (B) d (B) e (B) f (C) g (A) h (A) i (E) j (A) k (A)

A Fyrquel EHC 25.61 37.23 9.82


B Fyrquel EHC 3.49 4.19
C Fyrquel EHC 3.49 7.55
D Fyrquel EHC 7.55 14.6
E Fyrquel EHC 14.6 31.9
F Fyrquel EHC 3.53 7.29 2.88 4.15 3.78 8.15
G Shell IRIS DR 1.16 0.97 6.03 1.39
H Shell IRIS DR 0.97 6.03
I Fyrquel EHC 3.42 3.65 3.74
J Fyrquel EHC 3.42 8.48 5.96
K Fyrquel EHC 1.95 4.97 3.62
L Fyrquel EHC 9.51 15.17
M Fyrquel EHC 4.66 5.28
N Fyrquel EHC Plus 3.05 8.84
O Fyrquel EHC Plus 7.77 14.77
P Fyrquel 220 0 0.85
Q Houghton 1146XC 4.16 8.17
R Shell Turbo DR46 9.31 8.41
S Shell Turbo DR46 8.41 16.74
T Shell Turbo DR46 16.74 14.62
U Shell Turbo DR46 14.62 13.72
TABLE 2. (Continued)

Fluid Fluid Type Initial a (D) b (D) c (B) d (B) e (B) f (C) g (A) h (A) i (E) j (A) k (A)

V Shell Turbo DR46 9.31 4.87


W Shell Turbo DR46 4.87 9.16
X Shell Turbo DR46 9.31 18.25
Y Shell Turbo DR46 18.25 10.66
Z Shell Turbo DR46 9.31 8.71
AA Shell Turbo DR46 8.71 9.74
AB Shell Turbo DR46 3.79 4.04
AC Fyrquel EHC 2.05 4.85
AD Fyrquel EHC 4.85 7.3 10.27 7.9
AE Fyrquel GT 0 0

HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106


AF Fyrquel GT 0 0 0.74
AG Fyrquel EHC Plus 1.66 2.05 2.56 1.86
AH Fyrquel EHC Plus 2.05 4.71
AI Fyrquel EHC 1.97 1.86 2.68
AJ Gulf Turbofluid 46XC 2.21 2.5
AK Fyrquel EHC Plus 3.16 5.22 11.53
AL Fyrquel EHC Plus 11.53 8.85
AM Castrol Anvol PE 46 HR 3.53 7.9
AN Castrol Anvol PE 46 HR 7.9 18.4
AO Castrol Anvol PE 46 HR 4.57 4.53 5.47
AP Castrol Anvol PE 46 HR 2.57 6.9
AQ Castrol Anvol PE 46 HR 6.9 8.94

89
90
STP 1573 On Fire Resistant Fluids
TABLE 2. (Continued)

Fluid Fluid Type Initial a (D) b (D) c (B) d (B) e (B) f (C) g (A) h (A) i (E) j (A) k (A)

AR Fyrquel EHC N 1.95 5.92


AS Castrol Anvol PE 46 HR 5.62 7.62 7.97
AT Fyrquel EHC 0.67 16.64
AU Reolube Turbofluid 46XC 6.47 8.21 14.78
AV Fyrquel EHC 4.39 6.12
AW Reolube Turbofluid 46XC 1.76 3.44
AX Fyrquel EHC 7.78 23.92
HOBBS AND DUFRESNE, DOI 10.1520/STP157320130106 91

References

[1] Phillips, W. D., “Phosphate Esters,” Handbook of Hydraulic Fluid Technology, G. E. Totten,
Ed., Marcel Dekker, NY, 2000, pp. 1025–1094.

[2] Lin, C. R., “What is Streaming Current-Driven Corrosion?” FRH Journal, Vol. 2(1), 1981,
pp. 17–21.

[3] ASTM D1169-11: Standard Test Method for Specific Resistance (Resistivity) of Electrical
Insulating Liquids, Annual Book of ASTM Standards, ASTM International, West Consho-
hocken, PA, 2011.

[4] Staniewski, J. W. G., “Maintenance Practices for Steam Turbine Control Fire Resistant
Fluids: Part 1,” J. Synth. Lubr., Vol. 23, 2006, pp. 109–121.

[5] Staniewski, J. W. G., “Maintenance Practices for Steam Turbine Control Fire Resistant
Fluids: Part 2,” J. Synth. Lubr., Vol. 23, 2006, pp. 123–135.

[6] EPRI, Electrohydraulic Control (EHC) Fluid Maintenance Guide, EPRI, Pala Alto, CA,
2002.

[7] Kimble Chase, “Gravity Chromatography Columns With Fritted Disc and PTFE Stopcock
Plug—Product Specification Sheet,” Kontes Article No. 420540-0245, Kimble Chase,
Vineland, NJ.

[8] Boury, B., Nair, R. G., Samdarshi, S. K., and Makiabadi, T., “Non-hydrolytic Syn-
thesis of Hierarchical TiO2 Nanostructures Using Natural Cellulosic Materials as
Both Oxygen Donors and Templates,” New J. Chem., Vol. 36, 2012, pp.
2196–2200.

[9] Handbook of Mineralogy, J. W. Anthony, R. A. Bideaux, K. W. Bladh and M. C. Nichols,


Eds., Mineralogical Society of America, Chantilly, VA, 2013.

[10] Izci, E. and Izci, A., “Dielectric Behavior of the Catalyst Zeolite NaY,” Turk. J. Chem., Vol.
31, 2007, pp. 523–530.

[11] ASTM D664-07: Standard Test Method for Acid Number of Petroleum Products by
Potentiometric Titration, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2007.

[12] ASTM D5185-09: Standard Test Method for Determination of Additive Elements, Wear
Metals, and Contaminants in Used Lubricating Oils and Determination of Selected Ele-
ments in Base Oils by Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-
AES), Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2009.

[13] ASTM D4929-07: Standard Test Method for Determination of Organic Chloride Content
in Crude Oil, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2007.

[14] ASTM D7546-09: Standard Test Method for Determination of Moisture in New
and In-service Lubricating Oils and Additives by Relative Humidity Sensor, An-
nual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2009.
92 STP 1573 On Fire Resistant Fluids

[15] Phillips, W. D., “A Comparison of Fire-resistant Hydraulic Fluids for Hazardous Industrial
Environments. Part II. Stability, Fluid Life and Environment,” J. Synth. Lubr., Vol. 14, 1998,
pp. 307–330.

[16] Phillips, W. D., “The Electrochemical Erosion of Servo Valves by Phosphate Ester Fire-
resistant Hydraulic Fluids,” Lubr. Eng., Vol. 44(9), 1987, pp. 758–767.

[17] Archie, G. E., “The Electrical Resistivity Log as an Aid in Determining Some Reservoir
Characteristics,” Trans. Am. Inst. Min., Metall. Pet. Eng., Vol. 146, 1942, pp. 54–62.
FIRE RESISTANT FLUIDS 93

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130119

Peter T. Dufresne1

Thirty-Seven Years of Fleet


Operating and Maintenance
Experience Using Phosphate
Ester Fluids for Bearing
Lubrication in Gas-Turbine/
Turbo-Compressor Applications
Reference
Dufresne, Peter T., “Thirty-Seven Years of Fleet Operating and Maintenance Experience Using
Phosphate Ester Fluids for Bearing Lubrication in Gas-Turbine/Turbo-Compressor
Applications,” Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 93–108, doi:10.1520/
STP157320130119, ASTM International, West Conshohocken, PA 2014.2

ABSTRACT
TransCanada operates one of the most sophisticated pipeline systems in the
world with a network of approximately 57 000 km (35 500 miles) of wholly
owned and 11 500 km (7000 miles) of partially owned natural gas pipeline, which
connects virtually every major natural gas supply basin/market and transports
20 % of the natural gas consumed in North America. This pipeline system
operates a fleet of several hundred gas-turbine/turbo-compressor packages,
which, in a majority of cases, use phosphate-ester-based lubricants for bearing
lubrication. Based on the number of installations, TransCanada is one of the
largest users of phosphate-ester lubricants in the world. Phosphate esters have
been used in this application because of their excellent fire-resistant and
lubrication properties. Mineral-oil-based lubricants, alternatively, used in this
application contain rust and oxidation inhibitors. These additives deplete over
time, which is one factor that limits their useful operating life. Phosphate-ester
lubricants do not generally contain additives and when used in conjunction with
effective condition monitoring and maintenance, can have an operating life >20
years. This extraordinarily long operating life offers significant environmental

Manuscript received July 17, 2013; accepted for publication January 21, 2014; published online April 4, 2014.
1
EPT Research, Calgary, AB T2B3R4, Canada.
2
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
94 STP 1573 On Fire Resistant Fluids

benefits. Since 1958, TransCanada has accumulated over 30  106 h operating


experience with phosphate-ester lubricants and has developed an extensive fluid
management and conditioning program to ensure safe, reliable, and cost
effective operation. A historical review of the condition-based fluid monitoring
and maintenance programs will be presented with a focus on the evolution of
these programs into their current modern and sophisticated forms. The
associated reduced environmental footprint and cost savings will be detailed.

Keywords
phosphate ester, Fuller’s Earth, ion exchange, lubricant maintenance, economic
analysis, environmental waste reduction

Introduction
Phosphate-ester lubricants have been used consistently at TransCanada since 1958
(P. E. Dufresne, private communication). These lubricants are used for bearing
lubrication in TransCanada’s industrial gas turbines and natural gas compressors.
A critical design requirement of natural gas compressors operating on both high-
pressure seal-oil (HPSO) systems and newer dry-gas seals is that during an emer-
gency shutdown as the result of a seal failure, the high-pressure oil pumps cannot
stop until all of the gas has been evacuated from the compressor. This typically
takes up to 6 min. Without the high-pressure seal-oil pumps operating during such
a shutdown, natural gas will flood the compressor station. As a result, the risk of
fire extends well beyond turbine shutdown. To protect against this risk, the lubri-
cant in use has to have self-extinguishing fire characteristics. The use of phosphate-
ester fluids provides TransCanada with the maximum level of safety and protection
against fire and has played an important part in the company’s ability to design the
overall pipeline system for remote and unmanned operation.
TransCanada uses two major brands of phosphate-ester fluid including both
natural and synthetic versions. TransCanada’s phosphate-ester lubricant-maintenance
program for the first 34 years was difficult, expensive, and time consuming. Despite the
lubricant’s excellent safety characteristics, its perceived extensive maintenance require-
ments and the resulting mechanical problems associated with poor lubricant condition
gave these fluids a very bad reputation. Thankfully, improved filtration technology and
maintenance practices were developed as part of an overall strategy at TransCanada to
optimize all maintenance and safety systems. Over the past 21 years, these improved
maintenance practices have not only resolved the historical issues, but have demon-
strated that phosphate-ester lubricants are safe, cost effective, and relatively easy to
maintain. This paper will review TransCanada’s historical maintenance practices during
this period and how these evolved into the sophisticated program currently in place.
There have been a number of major expansions of the TransCanada network
since 1957, which incrementally expanded the number of turbine compressor pack-
ages to over 270 units (including 22 electric drives and 244 gas turbines) by 1999.
DUFRESNE, DOI 10.1520/STP157320130119 95

The largest expansion occurred between 1990 and 1999 where compression
capacity was increased by 70 % to 5760 MW in total.

Mechanical Systems Background and Typical


Failure Pathways (Pre-1992)
In the common configuration that applies over 90 % of the time, there is either an
industrial turbine or aero-derivative turbine driving a power turbine compressor
package. In the industrial turbine packages, the turbine and power turbine share a
common PE lubricant reservoir. In the aero-derivative units, these systems use jet
engine lubricant, whereas the power turbine and compressor package use phosphate
ester. To understand the failure pathways and the costs associated with lubricant
failure, we will review the primary mechanical systems involved.

NATURAL GAS COMPRESSORS


All OEM’s design compressor rotors have one drive-end tilt-pad journal bearing,
one impeller-end tilt-pad journal bearing, one thrust bearing, and one HPSO as-
sembly (which behave like a standard journal bearing but with a very small oil flow
across the seal). Seal clearance is 0.001500 and is designed this way to minimize oil
flow across the seal during full pressure operation (1150 psi). The small amount of
oil flow across the seal amounts to about (3/4) of a barrel per month during normal
operations. The oil flow across the seal is required to maintain the seal temperature
below 150 F.
Over time, as the phosphate-ester lubricant degrades and accumulates degrada-
tion by-products, the lubricant will form insoluble deposits. These insolubles first
start to deposit on the high-pressure seal surface because of the minimal clearance
between the compressor shaft and the seal. As deposition continues, the first sign of
a problem is an increase in seal oil return temperature. There is no cure for this
problem, and once seal surface temperature exceeds approximately 220 F, the bab-
bitt seal surface melts and total seal failure occurs.
Upon total failure of the high-pressure seal oil (HPSO), the compressor control
system initiates a unit emergency shutdown. The fuel valve to the turbine closes
and the turbine spools reduce in speed until stopped. The internal volume of natu-
ral gas (20 m3) in addition to that in the compressor yard piping is evacuated to the
atmosphere (J. Carlson, private communication).
The HPSO pump(s) continue(s) to run until pressure within the gas compres-
sor falls below 25 psi. The time required to accomplish this total gas evacuation is
about 6 min. Historically, the most common compressor failure was HPSO related,
however, there were also frequent journal bearing and thrust bearing failures. Fail-
ure mode in these instances was the same as seen in the HPSO system; insoluble
degradation by-products deposited on the tilt-pad journal bearings or thrust-
bearing pads cause high-bearing temperatures until the babbitt melts causing total
bearing failure. Journal-bearing failures usually required replacement of the
96 STP 1573 On Fire Resistant Fluids

compressor rotor shaft, along with replacement of all compressor aerodynamic


rotating and stationary assemblies. Typical costs were about 80 % of new replace-
ment cost and could be up to $180,000 (P. E. Dufresne, private communication).
The corresponding unscheduled outage was typically about 3.5 days.

INDUSTRIAL TURBINE JOURNAL BEARINGS AND/OR THRUST BEARING


TransCanada’s industrial turbines generally have a forward tilt-pad journal bearing
(axial compressor forward tilt-pad bearing), a center bearing located within the
core of the combustor assembly, and a rear tilt-pad bearing located on the aft end
of the gas producer turbine rotating assembly. There is also a thrust-pad bearing,
generally located on the front of the gas-producer turbine rotating assembly. Bear-
ing clearances are typically larger than those in the natural gas compressors dis-
cussed above (about 0.006500 ). As a result, the buildup of insoluble PE degradation
products on the bearing surfaces involved requires more time than in the gas com-
pressor HPSO systems. Secondary damage from initial bearing failure was almost
always extensive, resulting in total failure of all rotating gas-path blading, as well as
total failure of non-rotating guide vanes or nozzles of the gas producer turbine as-
sembly. There were four to five gas producer turbine-bearing failures annually (P.
E. Dufresne, private communication). Damage costs were typically about 80 % of
the value of the initially purchased industrial turbine ($1.5–2.5  106) and the dura-
tion of the resulting unscheduled outage was never less than 3 weeks (P. E.
Dufresne, private communication).

POWER TURBINE THRUST BEARING


The power turbine is the low-pressure turbine associated with the gas-producer tur-
bine, but it is not mechanically connected to the gas-producer turbine. The power
turbine is the mechanical driver of the gas. The mode of failure in this application
was similar to the other bearing failures described above; however, almost all
thrust-bearing failures resulted in secondary damage caused by the resulting exces-
sive thrust-bearing clearance. This allows the high-pressure and low-pressure power
turbine rotors to contact the stationary blading of the power turbine. There was an
average of two power turbine thrust-bearing failures per year (P. E. Dufresne, pri-
vate communication). Failure costs were between $1.5–$2  106, with associated
unscheduled outages no less than 3 weeks (P. E. Dufresne, private communication).
A summary of the mechanical failures associated with lubricant deposits, the
costs to repair, the number of incidents per year, and the time required to repair is
summarized in Table 1 below. The resulting production losses per day are detailed
in Table 2 below.
In 1988 through 1994, production capacity was at >80 % meaning that any me-
chanical failure resulting in outage would frequently result in a production loss. In
1995, as new compressions came on line, the probability of loss was reduced to
50 %. After 1998, the probability of production loss was reduced to virtually 0 as
sufficient compression capacity existed throughout the network.
DUFRESNE, DOI 10.1520/STP157320130119 97

TABLE 1 Mechanical failures associated with lubricant deposits (1980–1992).

Mechanical Failure No. of No. of Days


Application Cost Per Incident Incidents Per Year to Repair Cost

Natural gas compressor/ 120 000 20 3.5 2 400 000


high-pressure seal assembly
Industrial turbine/thrust bearing 2 000 000 4 21 8 000 000
Power turbine/thrust bearing 1 500 000 2 21 3 000 000
13 400 000

In the following pages, an economic analysis is presented on lubricant con-


sumption, lubricant maintenance costs, and waste generated during different time
periods where different types of filtration methods were used. To present these
costs, certain variables and assumptions are listed in Table 3 below.

Historical Lubricant Maintenance Practices


1958–1985: LUBRICANT MAINTENANCE USING FULLER’S EARTH ACID-
REMOVAL FILTERS
From 1958 until 1985, the acid-removal systems used to treat TransCanada’s
phosphate-ester lubricants employed Fuller’s Earth filters. A large filter vessel hold-
ing six 1100  1900 Fuller’s Earth filters was connected in a kidney loop-type format
to the 1050-gal phosphate-ester reservoir. Each vessel contained approximately
297 lb of Fuller’s Earth per 12|550 lb of phosphate-ester fluid (2.4 %). The flow rate
through the fluid conditioning system was 360 GPH with a reservoir exchange rate
of one reservoir volume being filtered every 3 h.
During this period, the lubricant maintenance requirements for a given turbine
or compressor reservoir were related to which year of the lubricant life cycle each

TABLE 2 Production loss calculation associated with mechanical failures.

1000 Cubic ft/day 1988–1994 1995–1997 1998þ

Average production for each turbine/


compressor
Natural gas compressor/high-pressure seal 1 000 000
Industrial turbine/thrust bearing 1 000 000
Power turbine/thrust bearing 1 000 000
Toll charge $1.45 per GJ ¼ $1.0872 1.0872
per 1000 CF
Lost production associated per failure $1,087,200 $1,087,200 $1,087,200
Probability of production loss 80 % 50 % 0%
Total lost production value per day $869,760 $543,600 $0
98 STP 1573 On Fire Resistant Fluids

TABLE 3 Calculations, weights, and key parameters.

No. turbine/compressor packages 100


Reservoir volume (gal) 1050
No. years operating on phosphate-ester lubricants >37
Weight per Fuller’s Earth, alumina, or aluminosilicate filter (lb) 55
Weight of phosphate ester/gal 12
Types of acid filters used and average filter life
Average Fuller’s Earth filter life 1.5 months
Average alumina/aluminosilicate filter life 3.0 months
Average ion exchange filter life 18 months
1980–1992
Annual reservoir replacement 20 %
Fuller’s Earth filter disposal costs (per filter) 100
Man hours to replace a set of filters (hours) 4
Estimated labor rate including overhead (per hour) 50
Lubricant make-up volume per filter change (gal) 65

reservoir was in. As a result, fluid maintenance requirements varied directly with
the age of the fluid (see Table 3) [1]. Maintenance requirements and experiences
over this period are outlined below.
Until 1981, industry standard practice was to apply interrupted acid removal
only when acid number (AN) reached 0.50 mg KOH/g and then to discontinue acid
removal when AN reached 0.30 mg KOH/g. During the first year of operation on a
new lubricant reservoir, Fuller’s Earth cartridges were generally exhausted after two
operational cycles through the media over a 3-month period, and four sets of six (24)
cartridges were used during the first year. There was generally no fluid- or unit-
operating problems related to PE degradation during the first year of operation.
By the end of the second year of the lubricant life cycle, a set of Fuller’s Earth
cartridges was exhausted in approximately 6 weeks. As many as 48 cartridges were
used during the second year of fluid operation, and issues relating to fluid break-
down were observed for the first time. Metal soaps started to deposit on the surface
of high-pressure seals, reducing clearances and resulting in fluctuating seal oil pres-
sure. These factors contributed to fluid foaming along with elevated seal-operating
temperatures. Compressor seal mechanical failures because of lubricant deposits
became frequent (20 incidents per year). The cost of these failures and the resulting
production losses have been detailed in Tables 1 and 2.
Following the knowledge of best practices during this period, routine fluid anal-
ysis (AN, viscosity, water, and particulate) was unable to highlight these lubricant
problems. The importance of dissolved metals analysis, reported only from second-
ary lab analysis during this period, was not readily understood. It was not until a
later point that the sensitivity of the lubricant to dissolved calcium, magnesium,
and other metals would be clearly understood. Through the first 2 years of PE
DUFRESNE, DOI 10.1520/STP157320130119 99

lubricant service, 14 sets of Fuller’s Earth elements (84 cartridges) contributed to


calcium and magnesium levels between 50–90 ppm in the lubricant.
By the end of the third year of the lubricant life cycle, AN levels were >0.60 mg
KOH/g and acid-removal cartridges were exhausted in 3 weeks. AN could not be
lowered below 0.50 mg KOH/g. Calcium and magnesium levels became as high as
300 ppm. There were also signs of copper because of the corrosive effects of high
AN on copper core coolers. The impact of copper on lubricant breakdown became
apparent with significantly increased rates of acid production observed. High-
pressure seal failures continue to be a common occurrence. Evidence of more seri-
ous mechanical issues was finally observed with some industrial turbines experienc-
ing “hot” bearings.
By the end of the fourth year of the lubricant life cycle, AN levels have reached
as high as 1.9 mg KOH/g and Fuller’s earth filters had little to no effect. Mechanical
failures of industrial turbine and power turbine/compressor bearings from lubricant
deposits were frequent at six incidents per year.
Compressor seal failures, which were the most common, required 3.5 days to
repair, whereas compressor thrust bearing failures required 14 days to repair and
industrial turbines required a minimum of 21 days to repair (assuming that spares
were available). The associated production losses could, therefore, reach into the
tens of millions of dollars from a single turbine or compressor failure. TransCanada
had four spare industrial turbines that rotated in to replace failed units and at times
all four spares were in circulation. The cost of these common mechanical failure
pathways are estimated in Table 1.
It was clearly observed that the rate at which acid was produced increased
according to the acid number. Systems that had high acid numbers were almost
impossible to control as more acids were being produced than systems with low
acid numbers. For this reason, the use of interrupted filtration for acid removal was
discontinued in 1981 and full-time acid removal was adopted. No changes were
made to the physical properties of the lubricant maintenance system. Whereas con-
tinuous acid filtration was an improvement, the practicality of the situation changed
little as 4–12 sets of Fuller’s Earth elements (24–72 cartridges) continued to be used
on each reservoir, and by the end of the second year of service, the fluid AN was
always over 0.30 mg KOH/g. Clearly, one of the challenges associated with the use
of Fuller’s Earth was its limited AN-removal capacity. At TransCanada, one set of
Fuller’s Earth elements would reduce the AN by approximately 0.15.
From a turbine fleet perspective, 800 sets of Fuller’s Earth filters (4800 car-
tridges) were necessary each year, which made the labor and scheduling require-
ments enormous. One drum of new lubricant was required as make up for each set
of acid filters replaced, for a total of 800 drums per year. In addition, 20 % of the
turbine fleet required reservoir change-outs each year. Total lubricant maintenance
costs and lubricant purchased was approximately $5.1  106 (see Table 4). From a
waste perspective, significant user filter waste and lubricant waste was being gener-
ated that will be outlined later in this paper.
100 STP 1573 On Fire Resistant Fluids

TABLE 4 Summary of historical lubricant maintenance costs 1958–1985 using Fullers Earth.

Fluid Maintenance Costs


by Year (i.e., year 1 of
5-year Lubricant Life) Year 1 Year 2 Year 3 Year 4 Year 5 Total

No. of sets of acid filters 4 6 8 10 12 40


(six filters per set)
Filter costs 6,000 9,000 12,000 15,000 18,000 1,200,000
Disposal costs 2,400 3,600 4,800 6,000 7,200 480,000
($100 per filter)
Labor costs 800 1,200 1,600 2,000 2,400 160,000
(4 h at $50/h
per change)
Fluid make-up 11,000 16,500 22,000 27,500 33,000 2,200,000
costs from
filter change
(one drum per
filter change)
Fluid maintenance costs 20,200 30,300 40,400 50,500 60,600 4,040,000
No. of turbines 20 20 20 20 20
in each life cycle
Total annual maintenance 404,000 606,000 808,000 1,010,000 1,212,000 4,040,000
Reservoir replacement costs 1,050,000
Total maintenance 5,090,000
and lubricant
replacement costs

1984–1987: LUBRICANT MAINTENANCE USING ALUMINA ACID-REMOVAL


FILTERS
Fuller’s Earth’s inability to maintain acid number and prevent mechanical failures
prompted TransCanada to search for alternative acid filters with higher acid capacity.
In 1985, alumina-based acid-removal filters were tested and found to have higher acid-
removal capacity. At roughly twice the price of Fuller’s Earth filters, overall maintenance
costs per year remained similar, however. Initially, the alumina-based filters were able
to reduce AN by a wider margin compared to Fuller’s Earth (up to 0.25 mg KOH/g
reduction in AN per set); however, fluid-foaming issues started to develop on most res-
ervoirs within a year of installation. Foaming issues with the use of alumina are well
documented and have been attributed to the introduction of sodium into the PE fluid
from the alumina filters [2]. With the guidance of the fluid supplier, anti-foam additives
were added to problematic reservoirs. The benefit of using anti-foam additives was
short lived, however, and by the third and fourth years of the lubricant life cycle, the
use of additional anti-foam additives was ineffective. Using alumina filters, 3–10 sets of
filters were required per year and, by year 2, the AN was always over 0.30 mg KOH/g.
From a turbine fleet perspective, 600 sets (or 3600 cartridges) of alumina filters
were required each year, maintaining the strain on manpower requirements that
DUFRESNE, DOI 10.1520/STP157320130119 101

was present when Fuller’s Earth was used. One drum of new lubricant was required
to make up for each set of acid filters replaced, for a total of 600 drums per year. In
addition, 20 % of the turbine fleet required reservoir change-outs each year. Total
lubricant maintenance costs and lubricant purchases were approximately $4.7  106
per year (see Table 5).
While using alumina, turbine and compressor failures continued to occur on a
frequent basis, causing significant production losses as a result of reduced gas vol-
umes. During this period, production capacity continued to be operating at >90 %
throughout the pipeline, so any turbine or compressor failure had a 90 % probabil-
ity of resulting in production losses of up to 1.1  109 ft3 (BCF) per day, resulting in
up to $869,000 per day in lost revenue (see Table 2).

1987–1990: LUBRICANT MAINTENANCE USING ALUMINOSILICATE


ACID-REMOVAL FILTERS
In 1987, in an effort to resolve lubricant-foaming issues, low-sodium aluminosili-
cate-based acid-removal filters were tested. In most cases, these were found to sig-
nificantly reduce fluid foaming. Based on these results, and the severity of the
mechanical problems encountered while employing Fuller’s Earth and alumina as
acid filters, the fleet converted to the use of aluminosilicate filters within 1 year. Im-
mediate improvements were noted with the fluid-foaming problem and AN reduc-
tion was similar to the previous alumina-based acid-removal filters. Despite the
initial success, after approximately 6–12 months, lubricant gelling problems were
observed on a number of turbines and compressors. These gelling deposits were

TABLE 5 Summary of historical lubricant maintenance costs 1984–1987 using alumina.

Fluid Maintenance
Costs by Year (i.e., Year 1
of 5-Year Lubricant Life) Year 1 Year 2 Year 3 Year 4 Year 5 Total

No. of sets of acid filters 3 4 5 8 10 30


(six filters per set)
Filter costs 7,776 10,368 12,960 20,726 25,920 1,555,200
Disposal costs ($100 per filter) 1,800 2,400 3,000 4,800 6,000 360,000
Labor costs (4 h at $50/h 600 800 1,000 1,600 2,000 120,000
per change)
Fluid make-up costs from 8,250 11,000 13,750 22,000 27,500 1,650,000
filter change
(250 liters per change)
Fluid maintenance costs 18,426 24,568 30,710 49,136 61,420 3,685,200
No. of turbines in each life cycle 20 20 20 20 20
Total annual maintenance 368,520 491,360 614,200 982,720 1,228,400 3,685,200
Reservoir replacement costs 1,050,000
Total maintenance and 4,735,200
lubricant replacement costs
102 STP 1573 On Fire Resistant Fluids

noted throughout the lubricant system and particularly on mechanical surfaces


with lower temperatures. Whereas the gelling issue was obviously related to the use
of the new aluminosilicate filters, the mechanism of gel formation was not under-
stood at this time. Subsequent publication from the lubricant manufacturer advised
against using this form of acid-removal media on degraded fluid and recommended
these lubricant reservoirs to be replaced before use [3].
From a fleet perspective, 600 sets (or 3600 cartridges) of aluminosilicate filters
were required each year. Once again, the new filters did not address the ongoing
strain on manpower requirements. One drum of new lubricant was required to
make up for each set of acid filters replaced, for a total of 600 drums per year. In
addition, 20 % of the turbine fleet required reservoir change-outs per year. Total
lubricant maintenance costs and lubricant purchases were approximately 5.3  106
dollars per year (see Table 6). From a waste perspective, significantly used filter
waste and lubricant waste was still being generated. These tables will be discussed
later in this paper.
Turbine and compressor failures continued to occur on a frequent basis, which
were causing significant production losses as a result of reduced gas transmission.
From 1988 to 1994, system production remained very tight as >90 % as the pipeline
was operating at system capacity. Any turbine or compressor failure had a probabil-
ity of reducing gas flow by up to 1.1 BCF per day, costing up to $869,000 per day in
lost revenue.

TABLE 6 Historical lubricant maintenance costs 1987–1992 using alumniosilicate acid filters.

Fluid Maintenance Costs


by Year (i.e., Year 1 of
5-Year Lubricant Life) Year 1 Year 2 Year 3 Year 4 Year 5 Total

No. of sets of acid filters 3 4 5 8 10 30


(six filters per set)
Filter costs 10,800 14,400 18,000 28,800 36,000 2,160,000
Disposal costs 1,800 2,400 3,000 4,800 6,000 360,000
($100 per filter)
Labor costs (4 h at $50/ 600 800 1,000 1,600 2,000 160,000
h per change)
Fluid make-up costs from 8,250 11,000 13,750 22,000 27,500 1,650,000
filter change
(250 liters per change)
Fluid maintenance costs 21,450 28,600 35,750 57,200 71,500 4,290,000
No. of turbines in 20 20 20 20 20
each life cycle
Total annual maintenance 429,000 572,000 715,000 1,144,000 1,430,000 4,290,000
Reservoir replacement costs 1,050,000
Total maintenance and 5,340,000
lubricant replacement costs
DUFRESNE, DOI 10.1520/STP157320130119 103

1991–2013: LUBRICANT MAINTENANCE USING ICB ION EXCHANGE


By 1989, TransCanada’s usage of phosphate-ester lubricants came seriously into
question because of their inability to maintain the lubricant’s condition and, there-
fore, prevent costly turbine bearing and compressor seal failures. Whereas the
maintenance costs were very high, the underlying issue was the cost of the mechani-
cal failures and the resulting production losses. During this time, TransCanada was
planning a major system expansion and began to plan for the implementation of
new turbine compressor packages requiring the use of a fire-suppression system
because the economic viability of phosphate-ester lubricant use appeared to be
from a turbine fleet management perspective.
Between 1989 and 1991, with the support of one lubricant manufacturer, the
use of ion exchange resin purification was investigated. Earlier research from
Europe demonstrated the removal of acids using ion-exchange resins [4]. TransCa-
nada tested several ion-exchange resins from a number of manufacturers along
with several filter designs. It was subsequently determined that the use of certain
ion-exchange resins were effective for acid removal, but that there were a number
of technical issues and challenges to overcome.
By the end of 1991, the results from field trials showed that >90 % of reservoirs,
regardless of AN or fluid age could be returned to an AN of <0.10 mg KOH/g. It is
important to note that ion-exchange filters at this time were not available in the
marketplace, and, as a result, TransCanada had to design and contract out their
manufacturing. This development went through several prototypes before a top to
bottom flow configuration (axial flow) was found to be ideal. The use of axial flow
format filters was contrary to industry designs at this time, which used inexpensive
side flow configuration (radial flow). The use of axial flow presented challenges
because it resulted in higher differential pressures, which required more substantial
construction materials than conventional radial flow filters. It was also determined
that the use of stainless steel construction materials was required.
By 1992, the use of the optimized ICB ion-exchange filters developed was
expanded on a fleet-wide basis and the results exceeded expectations. The previous
issues of high AN, foaming and fluid deposits were eliminated. Lubricant maintenance
costs were reduced by 88 % to $677,000 per year (see Tables 7 and 8) but, more impor-
tantly, compressor seal failures and turbine-bearing failures were reduced to zero.

Cost Savings Associated with using Ion


Exchange (ICB) elements
FLUID-MAINTENANCE COSTS
If the ion-exchange lubricant-maintenance program was not implemented, the
lubricant maintenance costs at TransCanada after 1992 are estimated to have been
the same as the maintenance costs from the previous period 1987–1992 (Table 5).
Between 1992–2013, lubricant maintenance and purchasing costs were reduced by
104 STP 1573 On Fire Resistant Fluids

TABLE 7 Historical lubricant maintenance costs 1992–2013.

Fluid Maintenance Costs by Year


(i.e., Year 1 of 5-Year Lubricant Life) Year 1 Year 2 Year 3 Year 4 Year 5 Total

No. of sets of acid filters 0.7 0.7 0.7 0.7 0.7 3.3
(six filters per set)
Filter costs 3,492 3,492 3,492 3,492 3,492 349,200
Disposal costs 0 0 0 0 0 0
Labor costs 133 133 133 133 133 13,333
Fluid make-up costs 1,833 1,833 1,833 1,833 1,833 183,333
from filter change
Fluid maintenance costs 5,459 5,549 5,459 5,459 5,459 545,867
No. of turbines in 20 20 20 20 20
each life cycle
Total annual maintenance 109,173 109,173 109,173 109,173 109,173 545,867
Reservoir replacement costs 131,250
Total maintenance and 677,117
lubricant replacement costs

an estimated $4.6  106 per year for a total cost savings over 21 years of
$97,920,550. Whereas these cost reductions initially appear high, they are based on
one of the largest fleets of gas turbines in the world and represent a realistic
$1  106 cost savings per turbine over a 13-year period. It is interesting to note that
on a per turbine basis, very similar cost savings have been reported by other
phosphate-ester users from the power-generation industry where similar optimiza-
tions have occurred [5].

MECHANICAL COST REDUCTION


The nature of the mechanical failures associated with lubricant deposits were spe-
cific in nature and the costs associated with each specific failure is estimated in

TABLE 8 Lubricant maintenance program cost savings 1992–2013.

Fluid Maintenance Costs Cost Savings Total Cost


Savings 1992–2013 1987–1991 1992–2013 Per Year Years Reduction

Filter costs 2,160,000 349,200 1,810,800


Disposal costs ($100 per filter) 360,000 0 360,000
Labor costs (4 h at $50/h per change) 160,000 13,333 106,667
Fluid make-up costs from filter change 1,650,000 183,333 1,466,667
(250 liters per change)
Fluid maintenance costs 4,290,000 545,867 3,744,133
Reservoir replacement costs 1,050,000 131,250 918,750
Total maintenance and lubricant 5,340,000 677,117 4,662,883 21 97,920,550
replacement costs
DUFRESNE, DOI 10.1520/STP157320130119 105

Table 2. The HPSO system design was changed in 1997, which eliminated the risk
of seal failure associated with lubricant deposits. For this reason, HSPO cost savings
are only calculated up to 1997. Until 1997, HSPO failures cost $2.4  106 per year.
Whereas the rate of industrial turbine-bearing failures was relatively low (4 % of
fleet per year), the cost of failures was $8  106 per year, whereas power-turbine-
bearing failures cost $3  106 per year. In total, these mechanical failure costs were
$13.4  106 per year. If we assume that the turbine-bearing failures and power-tur-
bine-bearing failures would have occurred at the same rate moving forward, a cost
savings of $243  106 has been realized (see Table 9).

REDUCED PRODUCTION LOSSES


As discussed, turbine or compressor failures have historically had a high probability
of resulting in production losses. Until 1995, the probability of a production loss
was as high as 90 % as production capacity was constrained. After 1995, the proba-
bility of production loss was 50 % and, after 1997, sufficient compression capacity
existed that effectively reduced the probability of production loss to 0 %. Before
1995, TransCanada was having 26 mechanical failures per year, 90 % of which were
causing production losses of $860,000 per day.

ENVIRONMENTAL WASTE REDUCTION


The adoption of ICB ion exchange lubricant filtration at TransCanada has signifi-
cantly reduced filter and phosphate-ester lubricant waste. In Table 10, waste reduc-
tion calculations are outlined. The historical use of Fuller’s Earth filters required
4800 filters per year, whereas the use of use of alumina or aluminosilicate filters
required 3600 filters per year. Since 1992, the number of acid filters used has been
reduced to 402 filters, a reduction of 4400 Fuller’s Earth filters per year or 3200 alu-
minosilicate filters. The average reduction in number of acid filters used per year is
4267. The total reduction in filters disposed of is 89|607 units or 4.9  106 pounds
since the ion exchange filtration program began operation. With regard to lubricant

TABLE 9 Mechanical cost reductions achieved 1992–2013.

Mechanical Reduction in
Mechanical Cost Failure Cost No. of Annual No. of
Reductions Achieved Per Incident Incidents Per Year Total Years Total

Application component
Natural gas compressor/ 120,000 20 2,400,000 5 12,000,000
high pressure seal assembly
Industrial turbine/ 2,000,000 4 8,000,000 21 168,000,000
thrust bearing
Power turbine/thrust bearing 1,500,000 2 3,000,000 21 63,000,000
Total mechanical 13,400,000 243,000,000
cost reduction
106 STP 1573 On Fire Resistant Fluids

TABLE 10 Filter and phosphate-ester waste-reduction calculations 1992–2013.

No. of Fuller’s Earth filters that were used per year that would be landfilled until 1958–1988 4,800
No. of alumina or alumniosilicate filters that were used per year (1988–1992) 3,600
Average number of filters used per year 1958–1992 4,667
Ion exchange filters per year (1992–2013) 400
Reduced number of filters used per unit per year 4,267
Total reductions between 1992–2013 89,607
Total filter waste reduction (lb) 4,928,385
Annual phosphate-ester fluid consumption pre-1992 (gal) 65,000
Annual phosphate-ester fluid consumption post-1992 (gal) 8,800
Total reductions between 1992–2013 (gal) 1,108,200
Lubricant waste reduction (lb) 13,926,000
Total filter and lubricant waste reduction (lb) 18,854,385

consumption, total volume of phosphate-ester lubricant consumption has been


reduced from 65|000 gal per year to approximately 8800 representing a decrease of
86 % for a total reduction of 1.2  106 gal or 14  106 lb since 1992.

FINDINGS AND OBSERVATIONS


During the evolution of TransCanada’s PE fluid condition monitoring and mainte-
nance program, the following useful facts were discovered:
1. The rate of acid production is not linear and increases as acid number
increases. If AN values can be maintained at very low levels, less fluid mainte-
nance is required.
2. The presence of dissolved metals in a phosphate-ester lubricant makes them
less stable and prone to more rapid acid production.
3. Solid deposit formation in phosphate-ester applications occurs predominantly
when dissolved metals are present. When the metals can be removed and
maintained at very low levels, this primary pathway to solid deposit formation
can be avoided.
4. Fluid gelling associated was observed when the use of aluminosilicate filters
where applied to in-service degraded, PE lubricants.
As a result of these findings, the TransCanada lubricant maintenance and con-
ditioning program was optimized, best practices developed and written into stand-
ard TransCanada operation procedures (TOPS). These procedures require that all
turbine and compressor lubricants are tested each month by a third-party labora-
tory. The results are all trended by the centralized TransCanada laboratory. The
upper AN specification in the TOPS has been set at 0.07 mg KOH/g. When an AN
>0.07 mg KOH/g is reported, a re-test is requested. If the second sample AN is
>0.07 mg KOH/g, then an electronic work order is issued by the TransCanada lab-
oratory to replace the acid-removal filters. When this work order is assigned, a pre-
packaged set of replacement acid filters, particulate filters, gaskets, O-rings, and
used filter containers are dispatched from company inventory, and then
DUFRESNE, DOI 10.1520/STP157320130119 107

automatically re-ordered. In this manner, TransCanada has been able to achieve the
cost savings outline above.

Conclusions
TransCanada operates one of the largest fleets of gas turbines and compressors in
the world. Phosphate-ester lubricants are used in this application because they pro-
vide the highest level of safety and fire protection. TransCanada’s lubricant mainte-
nance has evolved into a highly sophisticated program with the strictest acid
number limits in the industry. This program was developed through extensive expe-
rience gained by resolving serious fluid-operating problems, which occurred until
1992.
TransCanada’s use of Fuller’s Earth (from 1958 to 1985), activated alumina
(from 1985 to 1987) and aluminosilicate (from 1987 to 1992) demonstrates that
these filtration media cannot maintain acid number effectively in turbine- or
compressor-bearing lubricant applications. Furthermore, these acid-filtration media
have been proven to contribute dissolved metals into the lubricant. The presence of
such metals significantly increases lubricant breakdown and creates secondary
chemical reactions from which deleterious solids are formed. These solids have a
tendency to deposit on mechanical surfaces causing failure, turbine outage, and lost
production.
Ion-exchange purification has been successfully used at TransCanada for the
last 21 years on over 100 gas turbines. The use of ion-exchange purification offers
significant advantages with regard to lubricant maintenance.
1. Ion-exchange filtration has very high acid-removal capacity.
2. As an engineered product, ion exchange does not contribute dissolved metals
or fine particulates.
3. Ion exchange will remove the metals contributed to the lubricant from other
types of acid-removal filters, reducing the propensity for the lubricant to create
damaging deposits, and eliminating the associated failures.
4. Ion-exchange filter life at TransCanada is approximately 18 months, which
significantly reduces manpower requirements and the associated lubricant
make-up required with filter changes.
The use of ion-exchange purification at TransCanada has resulted in significant
environmental savings. Filter waste has been reduced by 5  106 pounds, and lubri-
cant waste by 14  106 pounds for a total waste reduction of 19  106 pounds. Sig-
nificant greenhouse gas reductions have also been achieved from the elimination of
lubricant-related turbine and compressor failures. These values have not been pre-
sented in this paper.
The economic cost savings at TransCanada have been very positive. The use of
ion-exchange filtration has reduced fluid maintenance costs by $4.6  106 per year
or $97  106 since program inception. This translates to an average cost savings of
$1 million per turbine since program inception, or $46,000 per turbine per year.
108 STP 1573 On Fire Resistant Fluids

The return on investment (ROI) of the fluid maintenance program is 689 % per
year excluding the reduction in mechanical failures.
The chronic mechanical failures that were previously occurring each year from
lubricant deposit issues have been completely eliminated. These historical failures
were previously costing $13.4  106 per year. Had the last 21 years of failures not
been prevented, the continued usage of phosphate-ester lubricants may have been
impractical from an economic standpoint. Fortunately, through the efforts of many,
TransCanada was able to develop and implement state-of-the-art lubricant condi-
tion monitoring and maintenance practices that have made the program the success
it is today.

References

[1] Dufresne, P. E., “14 Million Hours of Operational Experience on Phosphate Ester Fluids as a
Gas Turbine Main Bearing Lubricant,” 12th International Colloquium Tribology, Esslingen,
Germany, Jan 12–13, 2000.

[2] EPRI, “Electrohydraulic Control (EHC) Fluid Maintenance Guide,” Product ID 1004554,
Electric Power Research Institute, Pala Alto, CA, 2002.

[3] Nobel, Akzo, “Selexsorb GT Power Generation Application Guide,” T1058, Ludwigshafen,
Germany (Oct 1, 1999).

[4] Brandt, F. C. J., Manigley, G., and Trost, R., “A New Method of Purifying Synthetic Hydraulic
Fluids by Means of Ion Exchange,” Presentation on Synthetic Fluids as Hydraulic and
Lubricating Oils in Turbine Plants, Technical Academy Esslingen, Ostfildern, Germany, Jan-
uary, 1986.

[5] Elangovan, R., “Ion Exchange Resin Treatment of Phosphate Ester Fluids in Pulau Seraya
Power Station,” Internal Publication at Pulau Seraya Power Station, 2003.
FIRE RESISTANT FLUIDS 109

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130118

Patti Cusatis,1 John Sherman,2 Paul Fasano,3


and Roland Bishop4

Property and Performance


Evaluation of Water Glycol
Hydraulic Fluids
Reference
Cusatis, Patti, Sherman, John, Fasano, Paul, and Bishop, Roland, “Property and Performance
Evaluation of Water Glycol Hydraulic Fluids,” Fire Resistant Fluids, STP 1573, John Sherman,
Ed., pp. 109–125, doi:10.1520/STP157320130118, ASTM International, West Conshohocken, PA
2014.5

ABSTRACT
Water glycol hydraulic fluids are the most used fire-resistant hydraulic fluids in
the world. A selection of commercial water glycol hydraulic fluids were chosen
based on their individual chemistry and being representative of the scope of
water glycol hydraulic fluid types in use within the general manufacturing
industry today. These fluids were then measured by a series of property and
performance tests to evaluate key fluid parameters, including pump
performance, lubricity, corrosion protection, and wear performance. The results
are discussed, with recommendations as to their suitability for use in general
industrial applications.

Keywords
water glycol hydraulic fluids, performance properties, fire resistant

Manuscript received July 17, 2013; accepted for publication October 23, 2013; published online April 4, 2014.
1
BASF Corporation, 500 White Plains Rd., Tarrytown, NY 10591, United States of
America, e-mail: patti.cusatis@basf.com
2
BASF Corporation, 1609 Biddle Ave., Wyandotte, MI 48192, United States of
America, e-mail: john.sherman@basf.com
3
BASF Corporation, 500 White Plains Rd., Tarrytown, NY 10591, United States of
America, e-mail: paul.fasano@basf.com
4
R. J. Bishop and Associates, Inc., 32 Donald Ln., Ossining, NY 10562, United States of America,
e-mail: rjbishopjr@aol.com
5
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
110 STP 1573 On Fire Resistant Fluids

Introduction
Water glycol hydraulic fluids are one class of fire-resistant hydraulic fluids denoted as
HFC as per ISO 6743/4 [1]. These fluids are composed primarily of water, glycol, and a
thickener. The water (typically in the range of 35 %–45 %), provides the property of fire
resistance to the fluid and the glycol component; ethylene glycol, diethylene glycol, or
propylene glycol give the fluid its low-temperature characteristics and additional viscos-
ity. Another key component is a high-molecular-weight, water-soluble polyalkylene gly-
col thickener, which provides the lubricity and viscosity properties. Various additives
provide other necessary attributes, including corrosion inhibition, anti-wear perform-
ance, and antifoam properties. These fluids are particularly suitable for demanding hy-
draulic applications where there is a high fire risk in medium-pressure hydraulic systems
(1500–3500 psi), such as those in the steel-making and die-casting industries [2].
In addition to fire resistance, these fluids have several advantages as compared
to mineral oil hydraulic fluids; they do not form sludge, they have low pour points,
and high viscosity indices. The disadvantages are that they can only be run at rela-
tively low operating temperatures (<65 C) and are incompatible with some metals,
plastics, and elastomers. In some cases, they form thinner hydrodynamic films as
compared to mineral oils of equivalent viscosity. Low fatigue life, high wear rates,
and cavitation erosion can be issues with water glycol fluids when used in hydraulic
systems for which they are not suited [2].
Several commercial water glycol hydraulic fluids have been evaluated for a wide
range of properties, including viscometric characteristics, corrosion and wear prop-
erties, and vane-pump efficiency. These properties will be reviewed here with the
intent to show the variation in the products available today. A common difficulty in
developing and evaluating lubricants is the identification and application of appro-
priate screening tests. It is frequently advantageous to utilize a range of performance
variables in the characterization of a fluid. Additionally, the presence of water in
these fluids poses some challenges from the standard test procedures typically used
for mineral-oil-based hydraulic fluid testing.
The fire resistance and environmental aspects of the fluids were not evaluated
here. Nonetheless, based on the individual technical bulletins of the fluids and
knowledge of their composition, all the fluids would be FM approved under FM
Approval 6930 [3], and would meet the fire-resistance requirements of ISO 12922
[4]. They would be a grade B or C according to the ISO 15029-2 Spray Test—Stabi-
lized Flame Heat Release Method [5]. Also, by their chemistry and composition,
many of the water glycol hydraulic fluids could pass biodegradability tests (such as
OECD 301) [6] and aquatic toxicity tests (such as OECD 203) [7].

Discussion
FLUID SELECTION
The water glycol hydraulic fluids evaluated in this paper were chosen from the
types known as standard water glycol fluids (>40 % water) and those known as
CUSATIS ET AL., DOI 10.1520/STP157320130118 111

high-performance water glycol fluids (<38 % water) as a representative sampling


of those on the market. The standard water glycol fluids may further be defined
as fluids typically used in piston pumps at hydraulic pressures no greater than
3500 psi, whereas high-performance water glycol hydraulic fluids may be used sat-
isfactorily in some piston pumps with operating pressures of 5000 psi. The sample
set tested includes two products classified as high performance and six products
classified as standard water glycol hydraulic fluids.

PHYSICAL PROPERTIES
Table 1 is a summary of some characteristic properties of the eight water glycol hy-
draulic fluids tested. In addition to a numerical designation, an extension is given to
each sample ID to give further information about a specific property of the sample
(some of which will be discussed in later sections). The designations are defined as
follows: ST, standard hydraulic fluid; HP, high-performance hydraulic fluid; LRA,
low reserve alkalinity; LMW, low molecular weight (thickener); AT, atypical
(thickener).
In terms of the reserve alkalinity, tested according to ASTM D1121 [8], two flu-
ids would be considered to have a low reserve alkalinity (<15 ml), whereas the
others are in the recommended range of at least 17 ml. The pH values, obtained
according to ASTM E70 [9], are all in the typical range for this type of fluid. Typi-
cally, the water content is between 35 % and 45 % [2] and this is the case in each of
the materials tested here. A review of the impact of the differences in these proper-
ties will be discussed in further detail in the following sections.

ANALYTICAL CHARACTERIZATION
Each product was analyzed by nuclear magnetic resonance (NMR) spectroscopy
to determine the ratio between ethylene oxide and propylene oxide monomer
amounts in the polyalkylene glycol thickener portion of the water glycol

TABLE 1 Summary of physical properties of the water glycol hydraulic fluids.

Reserve Alkalinity pH at 25 C Water Content Kinematic Viscosity


(ml) (%) at 40 C (mm2/s)

Sample ID ASTM D1121 [8] ASTM E70 [9] Karl Fischer ASTM D445 [10]

1-ST 18.1 9.7 40.5 42.50


2-HP 17.3 9.3 36.6 47.43
3-LRA 13.2 9.8 44.7 38.58
4-LRA 13.4 9.7 41.9 42.20
5-HP 19.8 9.4 35.5 47.80
6-ST 19.9 9.5 43.5 40.80
7-LMW 17.6 9.7 42.5 41.90
8-AT 22.6 9.7 41.5 42.44
112 STP 1573 On Fire Resistant Fluids

formulation. The fluids were dried under a heat lamp to remove the volatile com-
ponents and the viscous residues were analyzed by 1HNMR and 13CNMR using a
Bruker AVANCE II 400. For each test fluid, the monomer ratio of the polyalky-
lene glycol thickener was determined to be 75 % ethylene oxide and 25 % propyl-
ene oxide.
Each sample was analyzed by Fourier Transform Infrared Spectroscopy (FTIR)
using a Thermo Scientific Nicolet 4700 to determine the type of thickener used.
The fluids were dried under a heat lamp to remove the volatile components and the
water, then the viscous residues, were analyzed using an attenuated total reflectance
cell. Each sample possessed a strong ether band in the region from 1050 to
1150 cm1, as well as absorbance in the hydroxyl region (3500 cm1) of the infra-
red spectrum. These features indicate that all the test fluids can be identified as hav-
ing polyether-type residual materials. There was some variation in the hydroxyl
region because of differences in the specific initiator compounds used resulting in
differences in the specific polyalkylene glycol polymer structure. There was also
variation in the carbonyl region (1600–1800 cm1) because of differences in the
additive chemistry used. An example of a spectrum of one of the test fluids is
shown in Fig. 1.
A subset of the products was tested for polymer distribution using a Waters
Alliance 2696 Separation Module Size Exclusion Chromatograph. The samples
were prepared in tetrahydrofuran (THF) containing 3 % deionized water. Duplicate
injections of each sample were made. The molecular weight values were determined
relative to a set of polystyrene standards. Figure 2 is an overlay of the

FIG. 1 Fourier transform infrared spectrum of one test fluid.


CUSATIS ET AL., DOI 10.1520/STP157320130118 113

FIG. 2 Size-exclusion chromatograms of samples 2, 3, 7, and 8.

chromatograms of four of the eight samples. As you move from left to right, the
molecular weight decreases. Samples 2 and 3 have similar molecular weight charac-
teristics with regard to molecular weight range and polydispersity. Sample 8 has a
much narrower polymer distribution than samples 2 and 3 with only a slight
amount of a lower molecular weight fraction. Sample 8 has a higher molecular
weight as compared to samples 2 and 3. Sample 7 has a significantly lower molecu-
lar weight than the other three samples and is mono-modal in nature.

VISCOMETRICS
As in any lubricant, viscosity and viscosity index are key properties for water
glycol hydraulic fluids. The kinematic viscosity values at 40 C were obtained for
all eight samples according to ASTM D445 [10] and shown in Table 1. Because
the water content of these fluids is quite high, it has a significant impact on the
fluid viscosity. The water content of each product was determined using Karl Fi-
scher titration methodology [11]. The water content for the products tested is in
the typical range from 35 % to slightly higher than 45 %.
The viscosity index (VI) is a measure of the relationship between the viscosity
and temperature of the fluid. The higher the viscosity index, the less the fluid vis-
cosity changes with fluid temperature. This parameter must be considered in deter-
mining the correct hydraulic fluid for a specific system operating temperature
range. The VI is an arbitrary number based on the measured viscosity at 40 C and
114 STP 1573 On Fire Resistant Fluids

FIG. 3 Three-temperature curve of kinematic viscosity.

100 C as described in ASTM D2270 [12]. Polyalkylene glycols thickeners have the
inherent advantage of high viscosity indexes [2].
As calculated in ASTM D2270, the viscosity index is obtained by measuring the
kinematic viscosities at 40 C and 100 C. Obtaining the 100 C kinematic viscosity
of water glycol hydraulic fluids is troublesome because of the high water content, so
a variation of the standard test method was used. Kinematic viscosity determina-
tions at 20 C, 40 C, and 80 C were performed and a 3-point line created (see
Fig. 3). Extrapolation of the line to 100 C was done to obtain that kinematic value.
Using the 100 C value determined through extrapolation and the measured

40 C viscosity, the viscosity indexes were calculated and are shown in Fig. 4. There
is a variation in the calculated viscosity index values among the eight samples. Sam-
ples 2 and 3 (pattern shading) are highlighted because they both have the same
thickener, but the sample with the higher water content has a better VI value. It
should also be noted that the other high-performance (lower water) product also
resulted in a lower VI result relative to the other samples in this set. This informa-
tion lends itself to further study regarding the incorporation of the thickener in the
formulations to determine how this impacts the viscometric properties of the fluid.

CORROSION PROPERTIES
Hydraulic fluids must be able to protect the entire hydraulic system against rust
and corrosion. Evidence of corrosion protection can be obtained by standard test
methods ASTM D665 [13] and ASTM D130 [14]. Ferrous corrosion testing by
ASTM D665 has two procedures: procedure A, where fresh water is added to the
test sample, and procedure B, where sea water is added to the test sample prior to
testing. ASTM D665 procedure A was completed adding no additional water to the
samples because the samples already contain significant levels of water. In ASTM
CUSATIS ET AL., DOI 10.1520/STP157320130118 115

FIG. 4 Viscosity index values.

D665 procedure B, synthetic sea water was added to the test fluid according to the
standard test procedure. Yellow metal corrosion testing was completed using the
ASTM D130 under standard conditions. It should be noted that, although not eval-
uated in this paper, vapor-phase corrosion properties of water glycol hydraulic flu-
ids are very important because of the presence of significant water. All water glycol
hydraulic fluids contain vapor-phase corrosion inhibitors, which are typically alkyl
alkanolamines or heterocyclic amines with appreciable vapor pressure. Test meth-
ods to evaluate vapor-phase corrosion protection include: ASTM D5534-
94(2011)e1: Standard Test Method for Vapor Phase Rust-Preventing Characteristics
of Hydraulic Fluids [15].
Table 2 is a summary of the corrosion testing performed on each sample. All
samples passed the ferrous corrosion testing according to ASTM D665A. All eight
samples passed the copper strip corrosion test, either with a rating of 1A or 1B.
In regard to procedure B for the ferrous corrosion test, ASTM D665B, one sam-
ple failed with a significant amount of rust, whereas two samples were borderline
passes. A borderline pass means that there were conflicting test results: some of the
test results indicated there was no rust formation and some of the tests indicated a
small area of corrosion.
The more severe test conditions in procedure B are required in most hydraulic
fluid specifications. Five of the fluids tested here were able to clearly pass the more
stringent conditions. The failure in the other three products may indicate that the
type and concentration of alkyl alkanolamines used in those formulations were not
sufficient to provide adequate corrosion protection under those conditions. This
should be kept in mind when the addition of water is made during operation. Water
glycol hydraulic fluids typically lose water during normal operation and water must
116 STP 1573 On Fire Resistant Fluids

TABLE 2 Summary of corrosion testing.

Ferrous Corrosion Ferrous Corrosion Copper Corrosion


(ASTM D665A) [13] (ASTM D665B) [13] (ASTM D130) [14], 100 C, 3 h

Sample ID 24 h, 60 C (No Water Added) 24 h, 60 C 3 h, 100 C

1-ST Pass Fail, 20 % rust 1A


2-HP Pass Borderline pass 1A
3-LRA Pass Pass 1A
4-LRA Pass Pass 1A
5-HP Pass Pass 1A
6-ST Pass Borderline pass 1A
7-LMW Pass Pass 1B
8-AT Pass Pass 1B

be added back into the system to maintain the viscosity and proper component
concentrations. Manufacturers of water glycol hydraulic fluids recommend that the
added water must be distilled, deionized, or have a specific maximum level of con-
ductance. One reason for this is to prevent corrosion as seen in this test, whereas
another is to prevent the creation of water insoluble salts that can clog filters. Metal
salts may also react with the fatty acid alkyl alkanolamine boundary lubricant used
in water glycol fluids to form a water insoluble fatty acid. Water used in production
or replenishment of water glycol fluids typically has a specific conductance of less
than 15 microsiemens per centimeter (lS/cm) [16].

WEAR PROPERTIES
There are increasing efforts to extend the life of machine equipment and a major
key to this is reducing the generation of friction. There are several bench tests that
measure wear properties. Attempts noted in the literature at correlating these types
of data with hydraulic equipment testing have been unsuccessful [17,18]. Four-ball
testing by ASTM D4172 [19] was performed on this data set. Because of the water
content, a variation of the test procedure was made. The test was performed at
55 C, 1800 rpm, 40-kg load for 60 min. These data, as shown in Fig. 5, indicate that
there is little variation in the wear scar and coefficient of friction results under these
test conditions. Therefore, very little information is provided in regard to differen-
tiation in wear properties using this test procedure. An alternative to the standard
bench tests will be presented in the next section on small vane-pump testing.

SMALL VANE-PUMP TESTING


A widely accepted method for measuring the wear potential for a hydraulic fluid is
the vane pump, known as the Vickers V-104C and now also performed with a re-
vised design called the Conestoga vane pump. The standard test, according to
ASTM D7043 [20] uses a 5-gal reservoir and runs for 100 h at a pressure of
2000 psi, a speed of 1200 rpm, with the fluid at a temperature of 65 C. The
CUSATIS ET AL., DOI 10.1520/STP157320130118 117

FIG. 5 Four-ball results for products.

difference in the weight of the ring and vanes before and after the test is used to
determine the wear. The typical limit for an approved fluid thirty years ago was
<100 mg total wear; however, improvements in water glycol hydraulic fluid formu-
lations has decreased the maximum wear limit for some OEM requirements to <50
mg total wear.
All eight samples were run on the vane pump using a “modified” ASTM D7043
method. The test conditions for the vane pump were as follows: 100 h of operation
at a pressure of 2000 psig, a speed of 1200 rpm, and a fluid temperature of 150  F
(65 C), but with a 3-l reservoir instead of the 5-gal reservoir specified in the ASTM
method. The 3-l reservoir will be more stressful on the fluids, thereby potentially
showing greater viscosity and wear distinction among the fluids.
Table 3 is a compilation of the four-ball results and the vane-pump wear results
sorted by increasing four-ball wear results. Figure 6 shows there is no correlation
between the four-ball wear scar and the four-ball coefficient of friction data. There is
also no correlation between the four-ball wear results and the small vane-pump wear
results, as shown in Fig. 7 (for better viewing, sample 4-LRA was not included in Fig. 7).
It should be noted that one of the high-performance products (5-HP) does have the
lowest vane-pump wear result, which is expected, but the vane-pump wear for the sec-
ond high-performance product (2-HP) is not distinguishable from some of the stand-
ard hydraulic fluid products. There is also no specific correlation between the water
content and either of the wear test methods. Figure 8 shows this lack of correlation
between the water content and the vane-pump wear (sample 4-LRA not included).
118 STP 1573 On Fire Resistant Fluids

TABLE 3 Four-ball and small vane-pump results.

Water Four-Ball Four-Ball Coefficient Ring Vane Total


Content (%) Wear (mm) of Friction Wear (mg) Wear (mg) Wear (mg)

6-ST 43.5 0.616 0.0922 24.9 6.8 31.7


2-HP 36.6 0.620 0.0693 23.0 4.7 27.7
5-HP 35.5 0.623 0.0544 4.5 5.0 9.5
1-ST 40.5 0.627 0.0727 8.9 3.3 12.2
4-LRA 41.9 0.628 0.0927 1010.0 12.5 1022.5
8-AT 41.5 0.629 0.0761 63.4 5.1 68.5
7-LMW 42.5 0.657 0.0646 10.7 7.3 18
3-LRA 44.7 — — 22.4 5.5 27.9

The standard test procedure ASTM D7043 only stipulates obtaining a wear
result, but this particular unit was configured to obtain hydraulic pump efficiency
also. The hydraulic pump efficiency (HPE) reported in this paper was measured by
calculating the ratio of pump horsepower/motor horsepower (PHP/MHP), where
PHP ¼ (flow, gpm  pressure, psig)/1714 and MHP was determined by three phase
power transducer attached to the electric motor that ran the V-104C pump [21].
By the suitable use of friction and viscosity modifiers, in hydraulic fluid formu-
lations, HPE can be increased, which will result in decreased operating costs.
Recently, in light of increased fuel costs, it has become an advantageous selling
point for fluid manufacturers to show efficiency data to the customer.
Table 4 contains the vane-pump test results along with the water content and kine-
matic viscosity sorted by decreasing vane-pump efficiency. The sample with the high-
est efficiency is one of the high-performance products having the lowest water content
and higher kinematic viscosity. The sample with the lowest efficiency (4-LRA)

FIG. 6 Four-ball coefficient of friction versus four-ball wear.


CUSATIS ET AL., DOI 10.1520/STP157320130118 119

FIG. 7 Vane-pump wear versus four-ball wear.

correlates with the sample that had a high wear value. It should be noted that the wear
value for this material was extremely high and the cause is being investigated. How-
ever, for the efficiencies from about 83 % to 93 %, it is difficult to see a definite correla-
tion between the vane-pump wear and vane-pump efficiency. This is shown in Fig. 9.

TRIBOLOGICAL TESTING
The vane-pump results are valuable, but the test is costly and time consuming. It would
be beneficial if a bench-test screening method could be developed. Ball-on-disk mea-
surement using the Mini Traction Machine (MTM) was pursued as a possible option.

FIG. 8 Vane-pump wear versus water content.


120 STP 1573 On Fire Resistant Fluids

TABLE 4 Pump test results with water content and kinematic viscosity.

Water Kinematic Viscosity Total Pump Total Pump


Content (%) (mm2/sec) Wear (mg) Efficiency (%)

5-HP 35.5 47.8 9.5 93.7


1-ST 40.5 42.5 12.2 93.2
6-ST 43.5 40.8 31.7 89.3
3-LRA 44.7 38.58 27.9 88.1
2-HP 36.6 47.43 27.7 87.7
7-LMW 42.5 41.9 18 87.4
8-AT 41.5 42.44 68.5 83.7
4-LRA 41.9 42.2 1022.5 80.2

This unit is designed to measure the traction characteristics of fluids. It consists of a


steel ball and steel disk driven independently to create mixed rolling and sliding con-
tact. These traction measurements are taken as a function of speed, temperature, load,
and the ratio between sliding and rolling, known as the slide–roll ratio (SRR). The
slide/roll ratio is defined as the difference in the speed of the two moving surfaces di-
vided by their average speed and multiplied by 100 %. As the sliding-to-rolling motion
increases, there is more stress on the fluid and the traction coefficient will increase. The
frictional force between the ball and disk is measured by a force transducer [22]. A
lower traction coefficient should correlate to less energy loss and, therefore, improved
energy efficiency [23]. This measurement is dependent on the fluid viscosity. Therefore,
valid results can only be obtained for materials that are the same viscosity.

FIG. 9 Vane-pump wear versus total vane-pump efficiency.


CUSATIS ET AL., DOI 10.1520/STP157320130118 121

For an example of the possible use of this technique, results for a subset of the
samples having similar viscosities are presented. Each sample was run on an MTM
using the ball and disk configuration. Smooth steel balls and disks were used. The
sample was placed in the test reservoir at a temperature of 55 C. Testing was per-
formed at four different slide–roll ratios (50, 100, 150, and 200) to determine the
impact of increasing severity. The load imparted was 35 N.
Figures 10 and 11 show the traction curves for samples 1, 8, and 9. The viscosity
results for these three samples are listed on the graphs to demonstrate that this fac-
tor is eliminated as a variable. Figure 10 shows that there are significant differences
in the traction coefficients between the three samples over the range of speeds tested.
At low speeds in the boundary region, there are the greatest differences between the
samples; as the speed increases, the variability between the samples decreases. For
these three samples, the sample having the lower traction coefficient had the highest
total pump efficiency. This same trend was obtained when the test was performed
using all the same conditions, but at a SRR of 150. This is shown in Fig. 11.

SHEAR STABILITY
Maintaining viscosity is an important factor in a hydraulic system. A high viscosity
index is required for proper operation, but ensuring that the viscosity is not altered
significantly over the service time is also imperative. The thickeners in water glycol
fluids could be susceptible to shearing; therefore, obtaining information about the
shear stability is of great interest.

FIG. 10 Traction curve of similar viscosity fluids (SRR ¼ 200).


122 STP 1573 On Fire Resistant Fluids

FIG. 11 Traction curve of similar viscosity fluids (SRR ¼ 150).

There are several different types of shear-stability test methods used for indus-
trial lubricants. The sonic shear test, according to ASTM D5621 [24], is typically
used for hydraulic oils. Another well-regarded shear testing method is the tapered
roller bearing shear test, also known as the KRL test according to CEC L-45 [25].
Because of the open system in both these tests and the elevated temperature gener-
ated over the duration of the test, these tests are not suitable for obtaining the shear
stability of water glycol fluids. There is a significant loss of water, which increases
the fluid viscosity making testing under these conditions meaningless for this type
of sample. Another test, ISO 20844; Petroleum and Related Products—

TABLE 5 Viscosity and water content change after small vane pump.

Percent Viscosity Change Percent Change in Water

1-ST 20.4 0.2


2-HP 18.1 2.5
3-LRA 10.6 1.8
4-LRA 18 0.5
5-HP 18.6 3.1
6-ST 3.7 0
7-LMW 6.2 1.2
8-AT 15 1.7
CUSATIS ET AL., DOI 10.1520/STP157320130118 123

Determination of the Shear Stability of Polymer-Containing Oils Using a Diesel In-


jector Nozzle [26], has been reported to have been used successfully to determine
shear stability of water glycol fluids with less than 1 % change in water content in
one instance and less than 2 % change in water with another sample. This is
another open system, so it may be presumed that special precautions on the part of
the specific test laboratory resulted in more valid results.
The use of the vane pump for shear testing was investigated. At the end of the
100-h test run, the end-of-test material was analyzed for both the kinematic viscos-
ity and the water content. Table 5 lists the results for the eight samples. All samples
show a viscosity decrease as would be expected under the test conditions of the
vane pump. The % water loss is the lowest we encountered in the alternative test
methods used. This should be considered as a valid test for obtaining shear data for
water glycol fluids.

Conclusions
Successful hydraulic equipment operation relies heavily on the performance of the
hydraulic fluid used in the system. Knowledge of both the physical properties and
performance characteristics is imperative in understanding whether a hydraulic
fluid will meet the needs of the hydraulic system. The most important properties
for testing water glycol hydraulic fluids are kinematic viscosity, pH, water content,
corrosion protection, shear stability, and wear characteristics.
Some standard tests for hydraulic fluids are not suitable or do not provide any
useful information when used to test water glycol hydraulic fluids. The eight com-
mercial water glycol hydraulic fluid samples tested here did pass all the standard
tests; however, the bench tests did not always reflect the claims found in the product
literature. This was exemplified in the wear and efficiency results for one of the
products designated as a high-performance product.
Alternatively, there are some tests that may be able to provide additional infor-
mation when modified for evaluation of water glycol hydraulic fluids. Atypical
methods for wear and shear stability were presented here, including the utilization
of the vane pump for providing shear stability and fluid efficiency data. The use of
the ball-on-disk apparatus shows promise for providing screening information
about the vane-pump efficiency of the fluid.
In general, correlation of bench tests to the pump wear and pump efficiency is
limited when one is reviewing the total pump efficiency alone. Efforts are underway
to separate the total efficiency into volumetric and mechanical efficiency, which
may provide a better way to distinguish hydraulic fluids.
In the commercial samples tested here, the use of some standardized tests per-
formed in a modified manner show promise for supplying valuable information.
This information, in combination with standard property testing, can be used when
selecting or recommending an appropriate fluid or when one is developing a new
or improved fluid.
124 STP 1573 On Fire Resistant Fluids

ACKNOWLEDGMENTS
The writers thank the following people for their testing expertise, data interpretation,
and meaningful ideas: Edmund Yee for the analytical work (BASF, Wyandotte, MI)
and Lee Ann Brugger, Glenn Phillips, and Jeffrey Schoonmaker for the application
testing (Tarrytown, NY).

References

[1] ISO 6743/4, 1999, “Lubricants, Industrial Oils and Related Products (Class L)—Classifica-
tion. Part 4: Family H (Hydraulic Systems),” The International Organization for Standard-
ization, Geneva, Switzerland.

[2] Rudnick, L., Synthetics, Mineral Oil and Bio-Based Lubricants, CRC, Boca Raton, FL,
2006, pp. 119–137, 551–555.

[3] FM Approval 6930, 2009, “Standard Flammability Classification of Industrial Fluids,” FM


Global, Johnston, RI.

[4] ISO 12922, 2012, “Lubricants, Industrial Oils and Related Products (Class L)—Family H
(Hydraulic Systems)—Specifications for Categories HFAE, HFAS, HFB, HFC, HFDR and
HFDU,” The International Organization for Standardization, Geneva, Switzerland.

[5] ISO 15029-2, 2012, “Petroleum and Related Products—Determination of Spray Ignition
Characteristics of Fire-Resistant Fluids. Part 2: Spray Test—Stabilized Flame Heat Release
Method,” The International Organization for Standardization, Geneva, Switzerland.

[6] OECD 301, 1992, “Ready Biodegradability,” The Organisation for Economic Co-
Operation and Development, Paris, France.

[7] OECD 203, 1992, “Fish, Acute Toxicity Test,” The Organisation for Economic Co-
Operation and Development, Paris, France.

[8] ASTM D1121-11: Standard Test Method for Reserve Alkalinity of Engine Coolants and Anti-
rusts, Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2012.

[9] ASTM E70-07: Standard Test Method for pH of Aqueous Solutions with the Glass Elec-
trode, Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2012.

[10] ASTM D445-12: Standard Test Method for Kinematic Viscosity of Transparent and Opa-
que Liquids, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2012.

[11] ASTM E203-08: Standard Test Method for Water Using Volumetric Karl Fischer Titration,
Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2012.

[12] ASTM D2270-10e1: Standard Test Practice for Calculating Viscosity Index from Kinematic
Viscosity at 40 and 100 C, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2012.

[13] ASTM D665-12: Standard Test Method for Rust-Preventing Characteristics of Inhibited
Mineral Oil in the Presence of Water, Annual Book of ASTM Standards, ASTM Interna-
tional, West Conshohocken, PA, 2012.
CUSATIS ET AL., DOI 10.1520/STP157320130118 125

[14] ASTM D130-12: Standard Test Method for the Corrosiveness to Copper from Petroleum
Products by Copper Strip Test, Annual Book of ASTM Standards, ASTM International,
West Conshohocken, PA, 2012.

[15] ASTM D5534-94e1: Standard Test Method for Vapor-Phase Rust-Preventing Characteris-
tics of Hydraulic Fluids, Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA, 2011.

[16] Bishop, R. J., “Maintenance and Analysis of Water-Glycol Hydraulic Fluids,” Mach. Lubr.,
Vol. 1, 2003, pp. 24–31.

[17] Totten, G. E., Bishop, R. J., and Kling, G. H., “Evaluation of Hydraulic Fluid Performance:
Correlation of Water-Glycol Fluid Performance by ASTM D2882 Vane Pump and Various
Bench Tests,” SAE Technical Paper Series 952156, SAE, New York, 1995.

[18] Totten, G. E., Bishop, R. J., and Kling, G. H., “Prediction of Hydraulic Fluid Performance:
Bench Test Modeling,” International Fluid Power Exposition and Technical Conference,
NFPA Technical Paper Series I96-2.5, NFPA, Quincy, MA, 1996.

[19] ASTM D4172-94: Standard Test Method for Wear Preventive Characteristics of Lubricat-
ing Fluid (Four-Ball Method), Annual Book of ASTM Standards, ASTM International,
West Conshohocken, PA, 2012.

[20] ASTM D7043-12: Standard Test Method for Indicating Wear Characteristics of Non-
Petroleum and Petroleum Hydraulic Fluids in a Constant Volume Vane Pump, Annual
Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2012.

[21] Fluid Power Designers’ Lightning Reference Handbook, 8th ed., Berendsen Fluid Power,
Piscataway, NJ.

[22] Mini Traction Machine (MTM) Test System; User’s Manual. (1999). PCS Instruments,
London.

[23] Sullivan, W., Oumar-Mahamat, H., Webster, M., and Brandes, E., “Lubricating Fluids With
Low Traction Characteristics,” U.S. Patent No. 7,732,389 (2010).

[24] ASTM D5621-07: Standard Test Method for Sonic Shear Stability of Hydraulic Fluids, An-
nual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2012.

[25] CEC-L-45-99, 1999, “Viscosity Shear Stability of Transmission Lubricants (Taper Roller
Bearing Rig),” The Coordinating European Council, Desford, U.K.

[26] ISO 20844, 2004, “Petroleum and Related Products—Determination of the Shear Stabil-
ity of Polymer-Containing Oils Using a Diesel Injector Nozzle,” The International Organi-
zation for Standardization, Geneva, Switzerland.
FIRE RESISTANT FLUIDS 126

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130103

Martin R. Greaves1 and Andrew Larson2

Anhydrous Fire-Resistant
Hydraulic Fluids Using
Polyalkylene Glycols
Reference
Greaves, Martin R. and Larson, Andrew, “Anhydrous Fire-Resistant Hydraulic Fluids Using
Polyalkylene Glycols,” Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 126–142,
doi:10.1520/STP157320130103, ASTM International, West Conshohocken, PA 2014.3

ABSTRACT
In industries such as steel- and aluminum-processing mining and aviation, fire-
resistant fluids are often an important consideration in choosing a hydraulic fluid.
Water-based hydraulic fluids are often an excellent choice when fire resistance
performance is the key selection criteria. However hydraulic pump reliability can
sometimes be a challenge, especially when equipment is operating under high
temperatures and pressures. Anhydrous hydraulic fluids are an alternative choice
for more severe operating conditions. Fire-resistant phosphate esters are known
to offer good fire resistance, but concerns about stability, toxicity, and cost has
led to innovation in the development of other anhydrous fire-resistant products
using synthetic esters, vegetable oils, and polyalkylene glycols (PAG). A
comparison of the performance of non-phosphorus-based anhydrous hydraulic
fluids including a synthetic ester, two conventional PAGs, and a new technology
based on an oil-soluble PAG (OSP) has been undertaken. In high-temperature
oxidation studies, PAGs and an OSP exhibited superior oxidation resistance
compared to a synthetic oleate ester. All fluids exhibited excellent hydraulic vane
pump wear performance but the ester showed high viscosity deterioration. An
unusual feature of the OSP fluid was its superior air release properties. This may
have practical benefits for hydraulic systems that use smaller reservoirs where

Manuscript received June 10, 2013; accepted for publication September 23, 2013; published online April 4,
2014.
1
The Dow Chemical Company, Bachtobelstrasse 3, CH-8810 Horgen,
Switzerland, e-mail: mrgreaves@dow.com
2
The Dow Chemical Company, 1691 N. Swede Rd., Midland, MI 48642, United States of
America, e-mail: arlarson@dow.com
3
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
GREAVES AND LARSON, DOI 10.1520/STP157320130103 127

cycle times are short. An examination of their fire-resistance properties using a


manifold test showed the ester fluid to have higher ignition temperatures but
lower ignitability factors than the PAGs and OSP fluids in the stabilized flame
heat release spray method. Conventional PAGs have been used as fire-resistant
fluids for over 20 years. Initial fire-resistance testing suggests that the new OSP
technology has similar performance and provides performance advantages such
as better low-temperature properties, faster air release, improved oxidation
stability, and compatibility with hydrocarbon fluids.

Keywords
polyalkylene glycol, oil soluble polyalkylene glycol, fire resistance, air release

Introduction
Petroleum-based fluids represent the vast majority of hydraulic fluids in operation
today in stationary and mobile equipment. Although these fluids often provide
adequate equipment performance, one recognized weakness is their poor fire-
resistance properties making them a less favorable choice than fire-resistant fluids
where there are environments with potential ignition sources. These include indus-
tries such as mining, aviation, steel processing, and aluminum die-casting because
hot metal, open flames, and sparks are all potential ignition sources if there is a leak
of hydraulic fluid [1–3].
Water-based hydraulic fluids offer an alternative choice to petroleum fluids
with the significant advantage of being much more fire resistant [4–6]. Research in
the 1940s led to the introduction of water glycol hydraulic fluids (now classified as
HF-C fluids according to ISO 6743-4) [7], which offered excellent fire-resistant
properties and for some hydraulic systems, good hydraulic pump performance.
These fluids were the first to use polyalkylene glycols (PAG) in fire-resistant fluids.
The PAG component in those formulations was a very high molecular weight poly-
mer (>10 000 Da) and used at treat levels of about 20 % by weight. Its primary use
was to provide elasto-hydrodynamic lubrication by thickening the base fluid, which
comprised water and a glycol (such as ethylene glycol). Because these fluids con-
tained high levels of water (>35 %) a further advantage was their relatively low
cost. Two practical disadvantages of water-based hydraulic fluids is that they need
to be carefully maintained in operation, and also they are limited for use in systems
where the bulk temperature is <55 C [8]. At higher temperatures, water has a high
vapor pressure, and this can lead to a rapid loss of water and, of course, a reduction
in its fire-resistant properties. These limitations subsequently led to the develop-
ment of anhydrous fire-resistant fluids that were safer to use than mineral oils but
could operate under more severe conditions than high-water-containing products.
Today anhydrous fluids and water-based fluids are commonly used in industries
where being fire resistant is a critical performance requirement.
Innovation over the last 40 years in the area of anhydrous fire-resistant fluids
has led to a number of new types of hydraulic products. These include fluids in
128 STP 1573 On Fire Resistant Fluids

TABLE 1 Classification of fire resistant fluids.

Classification Description

HFAE Oil-in-water emulsions containing >95 % water (v/v) and up to 5 % of additives


HFAS Chemical solutions in water containing >75 % water (v/v) and an additive package and
a water soluble polymer thickener
HFB Water-in-oil emulsions that typically contain >40 % water (v/v) and an additive package
HFC Water polymer solutions containing >35 % water (v/v), a high molecular weight
polyalkylene glycol thickener and an additive package
HFDR Anhydrous synthetic phosphate esters containing an additive package
HFDU Anhydrous synthetics (other than phosphate esters); examples include polyol esters,
vegetable oils, and polyalkylene glycols; the base oil represents about 95 % of the
formulation with the remainder being an additive package to enhance its performance

which the primary base oil is a phosphate ester, polyol ester, vegetable oil, or PAG
(also known as a polyether polyol). Their bulk fluid properties are enhanced with
components such as anti-oxidants and their surface properties are enhanced with
anti-wear and corrosion-inhibitor additive components. Each of these types of flu-
ids is classified according to ISO 12922 [9] and a summary of this is provided in
Table 1. For example, water glycol hydraulic fluids are classified as HFC fluids, phos-
phate esters as HFDR fluids, and polyol esters and PAGs as HFDU fluids.
Phosphate esters have been extensively researched and commercialized, and
today are still used in the aviation and power-generation industries [9–11]. Modern
fire-resistance tests generally show them to have superior fire resistance compared
to other common anhydrous fluids [12,13]. After more than 40 years of use, their
popularity seems to be declining in some applications because of environmental
and price concerns. Furthermore, it is known that phosphate esters are moisture
sensitive and prone to hydrolysis, which can lead to the formation of aggressive
acids that can be corrosive. In some applications, polyol esters and PAGs are pre-
ferred because they offer a better degree of fire protection than hydrocarbon oils
but are generally a lower cost option than phosphate esters. Polyol esters have been
marketed as fire-resistant fluids since the 1970s and PAGs since the 1990s. Both
have been successfully used in field operations. In recent years, new fire-resistant
hydraulic fluids based on vegetable oils have been promoted as well as formulations
containing blends of two or more base oils using natural vegetable oils, high oleic-
containing vegetable oils (e.g., high oleic canola oil and high oleic sunflower oil),
synthetic esters, and polyalkylene glycols as formulators seek to optimize the cost
versus performance balance [14]. One advantage of blending natural esters with
PAGs is improved deposit control. Under high-temperature conditions and in an
oxidative environment, natural esters oxidize primarily through the reaction of oxy-
gen with olefin moieties along the acid fraction of the triglyceride backbones. This
is known to lead to gums and sticky deposits. The inclusion of a PAG can help
minimize or even eliminate deposit formation by solubilization of oxidation
GREAVES AND LARSON, DOI 10.1520/STP157320130103 129

by-products. Also emerging are fluids marketed as fire resistant and biodegradable
and with a high level of renewable carbon content. This is becoming important in
applications where environmental considerations and fire resistance are key criteria
in a fluid-selection process.
Today there are four major groups of fire-resistant fluids that are recognized by
the International Organization for Standardization (ISO) and specification details of
these are defined in ISO 12922 and known as HFA, HFB, HFC, and HFD fluids. The
water-based fluids (HFA, B, and C) offer excellent fire-resistance results from existing
fire test standards. The HFD anhydrous fluids are generally known to have inferior
performance in common fire-resistance tests but nonetheless are considered to offer
improved performance over hydrocarbon oils. Fire-resistance performance is not
measured on the basis of one test alone, but an assessment is often made based on a
range of tests such as the spray flammability, hot manifold ignition, hot channel, wick
flame, and soaked cube test. Specific parameters such as fire point and flash point are
also relevant. Furthermore, there are many institutions that test and set standards for
fire-resistant fluids that include insurance companies and national governments. In
North America, FM Approvals, which is part of FM Global (formerly Factory Mu-
tual), is one of the world’s largest commercial and industrial property insurance
organizations, and has developed a certification program for the examination of fire-
resistance fluids. FM assesses fluids based on their spray flammability parameter
(SFP), and they are rated in two classes: “approved fluids” and “specification tested
fluids” [15]. Approved products provide a higher degree of fire protection and are
therefore more favorable. Certified fire-resistant fluids from FM Approvals are also
recognized in Asia and South America but not in Europe. Most water-based hydrau-
lic fluids are accredited with an “approved” rating, whereas anhydrous-based fluids
fall into both categories depending on the specific fluid.
In Europe, fire-testing protocols have long been established and follow the test-
ing requirements of the seventh Luxembourg report [16], which includes three fire-
resistance tests such as the Wick test, “pressurized spray” test, and “stabilized flame
heat release” test. The wide variety of fire-testing standards available has led to an
increased desire for tests to be standardized. The International Standards Organiza-
tion (ISO) is currently adopting a new global standard that essentially follows the
testing of the seventh Luxembourg report.
In this work, we examine the performance of non-phosphorus-based anhy-
drous fire-resistant hydraulic fluids and compare two conventional PAGs with a
leading synthetic ester-based fire-resistant fluid. Furthermore, a new class of PAG
has recently been developed known as oil soluble polyalkylene glycols (OSPs)
[17,18]. One objective of this paper is to examine the fire-resistance behavior of a
formulated fluid containing an OSP and compare this to other PAGs. The data that
will be presented is not extensive but will help provide direction on the suitability
of this class of fluid. Hydraulic fluids formulated with OSPs may provide an alterna-
tive option and choice to users of fire-resistant fluids. An assessment of their per-
formance and a comparison of their features and benefits will be undertaken.
130 STP 1573 On Fire Resistant Fluids

Anhydrous Fire-Resistant Hydraulic Fluids


for Evaluation
Four commercially available fluids were evaluated in this study and labeled as
fluids A to D. All the fluids are fully formulated fluids (i.e., they contain a
base oil and performance additive package) and are based on an ISO VG46
classification. Fluid A is a polyol oleate ester product that is known to be a
leading fire-resistant fluid in the steel-processing industry. Fluid B is a PAG
in which the PAG is a water-soluble type. Fluid C is also a PAG in which
the PAG is a water-insoluble type. Fluid D is an OSP in which the PAG is
an oil soluble type. Details of the fluids examined and their base oil chemis-
tries are shown in Table 2.

PAGs and OSPs and Their Chemistry


PAGs have been traditionally manufactured from the polymerization of ethylene
oxide (EO) or propylene oxide (PO) or a combination of both with a mono-

TABLE 2 Physical property measurements of fluids A to D.

Generic Base Fluid A Fluid B Fluid Fluid


Oil Chemistry (Ester) (PAG) C (PAG) D (OSP)

Trimethylol- PO
Test Propane EO/PO Homo- PO/BO
Base Oil Chemistry Method Trioleate Co-Polymer Polymer Co-Polymer

Kinematic viscosity ASTM D445 [19] 48.2 47.1 47.2 46


at 40  C (mm 2/s)
Kinematic viscosity ASTM D445 10.8 9.8 8.9 8.3
at 100  C (mm2/s)
Kinematic viscosity 320 390 480 460
at 0  C (mm2/s)
Viscosity index ASTM D2270 [20] 225 200 174 157
Pour point ( C) ASTM D97 [21] -37 -40 -45 -54
Flash point ( C) ASTM D92 [22] 278 276 274 267
Fire point ( C) ASTM D92 325 312 318 305
Density at ASTM D7042 [23] 0.92 1.03 0.99 0.96
15  C (g/ml)
Acid number ASTM D664 [24] 2.0 0.36 0.48 0.4
(mg KOH/g)
Vapor pressure ASTM E1719 [25] <0.01 <0.01 <0.01 <0.01
(mmHg at 25  C)
Water solubility at 5 % v/v Visual n/s s n/s n/s
Oil solubility at 5 % v/v Visual s n/s n/s s

Note: s, soluble; n/s, not soluble.


GREAVES AND LARSON, DOI 10.1520/STP157320130103 131

or polyhydric initiator in the presence of a base catalyst [26,27]. Initiators are


typically alcohols or polyols with labile hydrogens to facilitate alkoxylation.
When only one alkylene oxide is used, the resulting products are homo-
polymers. When two oxides are used, the polymers are co-polymers and can be
in the form of block, reverse block, or random co-polymers. Fluid B contains a
co-polymer of EO and PO and these polymers are often described as water-
soluble PAGs because many offer very high levels of water solubility. Fluid C is
a homo-polymer of PO. These polymers are often described as water-insoluble
PAGs. Product D is based on an OSP in which the polymer has been designed
from butylene oxide (BO) and PO feed stocks to produce a co-polymer. OSPs
have excellent oil solubility but very low water solubility. Water-soluble and
water-insoluble type PAGs have been successfully used as fire-resistant fluids
for over 20 years.

Physical Properties
Table 2 summarizes the physical properties of each product. Their water solubility
was tested by adding 5 % weight/weight (w/w) of each test fluid to de-ionized
water, mixing for 5 minutes at ambient temperature and visually assessing the
appearance of the fluid noting its sheening and miscibility characteristics. Simi-
larly their oil solubility was tested by adding 5 % w/w of each test fluid to an API
group I solvent neutral mineral oil (150SN) and visually assessing the appearance
of the fluid.

Modified Eaton (Vickers) V-104 C Hydraulic


Vane Pump Performance Testing
Hydraulic wear performance testing was conducted using an Eaton (Vickers) V-
104 C pump housing and Connestoga ring and vanes as explained in ASTM
D7043-12 (Standard Test Method for Indicating Wear Characteristics of Petro-
leum and Non-Petroleum Hydraulic Fluids in a Constant Volume Vane Pump)
[28]. However, a modified version of this procedure was used such that a reser-
voir capacity of 1 gallon was used instead of 5 gallons. Second, a comprehensive
cleaning procedure was employed before each test in which the equipment is
stripped down, cleaned, rebuilt, and flushed with new test fluid. In this way,
contamination from the fluid that was previously evaluated in the equipment is
minimized. Each hydraulic pump experiment was conducted at a bulk fluid tem-
perature of 65 C for 100 h, 1200 rpm, and a pressure of 2000 psi (pounds per
square inch), whereupon the weight loss of the vanes and the ring (cam) were
recorded and reported as total weight loss. The condition of the fluid before and
after each experiment was recorded by measuring its kinematic viscosity change
at 40 C and acid number change (AN). Table 3 summarizes the wear perform-
ance results.
132 STP 1573 On Fire Resistant Fluids

TABLE 3 Performance properties of fluids A to D.

Fluid A Fluid B Fluid C Fluid D


Base Oil Test (TMP (EO/PO (PO Homo- (PO/BO
Chemistry Method Trioleate) Co-Polymer) Polymer) Co-Polymer)

Ferrous corrosion ASTM D665A [30] Pass Pass Pass Pass

Copper corrosion ISO 2160 [31] 1b 1a 1b 1a


RPVOT (min) ASTM D2272 [32] 177 1145 1297 1050
Blown air oxidation
KV100 change after ASTM D2893B [33] 13.0 8.6 n/d 0.6
13 days (%)
TAN change after 2.0 0.4 0.14
13 days (mg KOH/g)
KV100 change after Modified 20.2 11.9 1.9
70 days (%)
TAN change after 2.5 1.5 0.16
70 days (mg KOH/g)
Air release at 50  C ASTM D3427 [34] 7 9 8 <1
(min)

Demulsibility, ASTM D1401 [35] 40/2038/2 Water Poor 41/2039/0


ml/ml/ml (min) soluble (20)

Four ball anti-wear ASTM D4172 [36] 0.36 0.55 0.50 0.31
(mm)
Eaton vane pump
V-104 C
Total weight loss of ASTM D7043 [28] 3.4 3.3 3.8 0.5
ring and vanes (mg)
Kinematic viscosity 14.7 0.5 1.4 1.0
at 40  C change (%)
TAN change 0 0.1 0.1 0.1
(mg KOH/g)

Note: n/d, not determined.

Eaton (Vickers) 35VQ25A Hydraulic Vane Pump


Performance Testing
Fluid D was also assessed in an Eaton 35VQ25A pump test using the method
described in ASTM D6973-05 in which an Eaton 35VQ25A vane pump is driven by
a 200HP motor [29]. The pumps inlet is supplied by a hydraulically pressurized
45 gal reservoir. The pump outlet is connected to an electronically controlled pres-
sure relief valve that is used to obtain the various pressures used throughout the
test. The pump operates at a speed of 2400 rpm, an outlet pressure of 205–210 bar
(3000 psi), inlet pressure of 0–2 psi, and an inlet temperature of 95 C. The
GREAVES AND LARSON, DOI 10.1520/STP157320130103 133

TABLE 4 Eaton 35VQ25A performance test on fluid D (OSP hydraulic fluid).

Result

Average cartridge wear (mg) 18


Viscosity change (%) 1.0
Water, before/after (%) 0.07/0.05
Acid number before/after (mg KOH/g) 0.19/0.24
Phosphorus, before/after (ppm) 137/117
Iron, before/after (ppm) <1/<1
Zinc, before/after (ppm) 2/6
Tin, before/after (ppm) <1/<1
Calcium, before/after (ppm) 6/7
Copper, before/after (ppm) <1/<1

duration of the test is 50 h for each cartridge. The evaluation requires a minimum
of three pump cartridges to be tested. The total weight loss of all the vanes should
be less than 15 mg. The total weight loss of cam (ring) from the individual cartridge
should be less than 75 mg. The total weight loss of the cam and vanes from each
cartridge tested should be less than 90 mg. Table 4 summarizes the test results.

Oxidation Testing
The oxidation performances of fluids A to D were measured in two ways. First, oxidation
measurements were made using the rotary pressure vessel oxidation test (RPVOT) as
described in ASTM D2272-11 (Standard Test Method for Oxidation Stability of Steam
Turbine Oils by Rotating Pressure Vessel) [32]. Second, oxidation performance was
assessed using experimental conditions described in ASTM D2893B-04 (Oxidation Char-
acteristics of Extreme Pressure Lubrication Oils) [33]. However, the data was reported in
two different ways. First, the change in kinematic viscosity at 100 C (KV100) was
reported after the 13-day test period (per the method) but also the change in acid num-
ber. Second, the test was extended beyond the usual 13-day test period to 70 days and
again the change in kinematic viscosity at 100 C and acid number were reported.

Four-Ball Wear
Four-ball wear measurements were made using ASTM D4172-94 (A Standard Test
Method for Wear Preventive Characteristics of Lubricating Fluid) (four-ball method)
[36]. The test evaluates the anti-wear properties of fluid lubricants in sliding contacts.

Air Release
Air release measurements on fluids A to D were made at 50 C using ASTM D3427-
12 (Standard Test Method for Air Release Properties of Petroleum Oils) [34]. Air
release measurements were also made on commercially available hydraulic fluids
134 STP 1573 On Fire Resistant Fluids

TABLE 5 Fire testing results.

RI valuea RI factor Manifold ignition


(ISO 15029-2) (ISO 15029-2) test (ISO 14935)

Phosphate ester (reference) 28 Class F >700


Fluid A 5 Class H 373–425
Fluid B 11 Class H 342–373
Fluid C 12 Class H 342–373
Fluid D 10 Class H 342–373
a
7th Luxembourg Report 3.1.3.6. The ignitability factor (RI) is graded into ranges with the least
flammable having an RI > 100.

including a water glycol fluid (HFC), a polyalphaolefin (PAO), a mineral oil, a satu-
rated polyol ester, and a vegetable oil all of which were ISO VG46 fluids from differ-
ent suppliers. The method measures the time for the entrained air content to fall to
a value of 0.2 % volume under a standardized set of test conditions and, hence, per-
mits the comparison of the ability of oils to separate entrained air under conditions
where a separation time is available. Entrained air is known to result in sponginess
and lack of sensitivity of the controls in turbine and hydraulic systems.

Fire-Resistance Testing
Fire-resistance testing was conducted using the stabilized flame heat release spray
method according to ISO 15029-2 [37]. The propane flow was 0.13 Nm3/h. Ignitability
indices (RI) are shown in Table 5 and Fig. 2.

Manifold Ignition Test (ISO 20823)


An International Standard (ISO 20823) [38] specifies a test method to determine
the relative flammability of fluids when in contact with a hot metal surface at a fixed
temperature, but it is also possible to gauge fluid ignition temperatures by adjust-
ment of the manifold temperature. It is primarily used to assess the resistance to
ignition of fire-resistant hydraulic fluids, which are, by definition, difficult to ignite.
A summary of the method follows. A 10-ml-test portion of fluid is dropped
from a predetermined height and at a specified rate, onto a tube heated to 700 C, or
in these experiments, at another defined temperature. The resulting spray is exam-
ined for flash or burn, both on the tube and after dripping from the tube. For each
fluid a temperature range is reported where flash or burning occurred. As a refer-
ence a commercially available fire-resistant phosphate ester fluid was used.

Results
Physical property measurements of fluids A to D are shown in Table 2 and perform-
ance property measurements in Table 3. The results of an Eaton 35VQ25A vane
GREAVES AND LARSON, DOI 10.1520/STP157320130103 135

pump test are shown in Table 4 for fluid D. Table 5 shows the fire-testing results for
fluids A to D and also a phosphate ester fluid that was used as a reference.

Discussion
In this work, we have examined the performance of three HFD chemistries (a syn-
thetic ester, two PAGs, and an OSP) in terms of their general laboratory perform-
ance in routine tests such as oxidation stability and vane pump testing as well as
their fire-resistance performance in a hot manifold test and stabilized flame heat
release method. In addition, we have explored their air release property, which is
becoming increasingly important to OEMs as equipment builders reduce weight
and size of some hydraulic circuitry.
Table 2 compares the physical property performance of fluids A to D. The pol-
yol oleate ester product (fluid A) exhibits the highest viscosity index (VI) of 225 but
also the highest pour point. Fluids B to D, which are all derived from polyether base
fluids, show a decrease in VI values from B to C to D. The primary reason for this
is that the alkoxide fraction of the polyether backbone has a higher carbon to oxy-
gen ratio for the OSP fluid (ratio is 3.5:1) than the PO homo-polymer (3:1) and the
EO/PO co-polymer (2.5:1). When the C/O ratio is 3.5 or above then higher degrees
of oil miscibility are possible [17] but the concession is a lower VI fluid. One of the
practical challenges perceived by end-users of conventional PAG hydraulic fluids
(such as fluids B and C) is that they are not oil miscible and therefore care must be
taken in good flushing practices when converting a hydraulic system from a hydro-
carbon oil to a PAG. When this occurs, PAGs usually provide excellent equipment
reliability performance and longevity. One benefit of the OSP technology is that
such extensive flushing procedures are not needed.
Fluid B is not oil soluble but is water soluble. This is a novel feature of hydraulic
fluids formulated with water-soluble polyalkylene glycols [39]. No other major type
of anhydrous-based hydraulic fluid has this feature. This can be an advantage when
they are used in hydraulic systems in which leaks to the environment and into
waterways occur, because the fluid does not form a sheen on the surface of water
and therefore does not lead to the residence of oil on water surfaces. Fluid B has
also a higher density than water and is readily biodegradable.
Each fluid showed excellent ferrous and copper corrosion protection. Their oxi-
dation stability was measured using two tests, which included the rotary pressure
vessel oxidation test (ASTM D2272) and a blown air oxidation test (ASTM
D2893B). In the RPVOT test, the ester fluid showed a low value of 177 minutes.
Fluids B to D showed significantly higher values and in the range 1050 to
1297 minutes. In the interpretation of the RPVOT data of PAGs, the authors have
observed in their laboratory, over a number of years, that the pressure decay curves
follow a very different pattern than that of hydrocarbon oils. Polyethers, including
fluids B to D, show a much steadier decay pattern, whereas hydrocarbon oils tend
to show a rapid pressure drop at a certain time. This suggests that the RPVOT data
136 STP 1573 On Fire Resistant Fluids

interpretation for polyethers should be different than hydrocarbon oils. This could
lead to higher RPVOT values being reported for PAGs. Further industry discussion
and exploration is needed in this area. In the blown air oxidation test, the viscosity
and acid number change of each fluid were measured after the 13-day test period
and also after an extended time of 70 days. The ester initially showed a decrease in
viscosity of 13 % after 13 days and then an increase to 20.2 % after 70 days. The
acid number change after 70 days was 2.5 mg KOH/g. Fluid B showed a lower level
of viscosity and acid number change after both 13 and 70 days. Fluid D showed
excellent stability. After 70 days, the fluid showed no significant evidence of aging
with a viscosity change of <2 % and an acid number change of <0.2 %. Therefore,
fluid D was more oxidatively stable.
Their demulsibility behavior showed various performances. Fluids A and D
exhibited good water-separation properties. Fluid B did not show any separation as
expected because this is a water-soluble fluid. Fluid C showed poor water
separation.
An Eaton vane pump test (ASTM D7043-12) using Connestoga ring (cam) and
vanes was used for assessing the wear properties of each hydraulic fluid. This test
measures the total weight loss of ring and vanes over a 100 h test period. All the flu-
ids showed a weight loss of <5 mg. This is considered to be an excellent result. Fluid
D exhibited a very low wear value of 0.5 mg and this was aligned with the very low
four-ball wear scar value of 0.31 mm using ASTM D4172. Surprisingly, fluid A
showed a significant amount of instability when its viscosity was measured post-
test. A viscosity change of 14.7 % was measured. This is unlikely to be caused by
oxidative degradation as the bulk fluid temperature was only 65 C and most likely
is because of shear degradation. Fluids B to D showed a negligible viscosity change
and therefore demonstrated excellent stability.
Fluid D was also assessed in an Eaton 35VQ25A pump test using the method
described in ASTM D6973-05. The average wear of three cartridges was 18 mg and
there was negligible change in the viscosity of the fluid when measured post-test as
shown in Table 4. In addition, no evidence of wear metals was observed. The reduc-
tion in the phosphorus level from 137 to 117 ppm was expected because this fluid
contains a phosphorus-containing anti-wear additive that is designed to coat metal
surfaces and provide an anti-wear protective film and therefore slowly deplete.
The air release properties were measured and are shown in Fig. 1. High air
release times were measured for fluids A to C. Most manufacturers of hydrocarbon-
based hydraulic fluids report values in the range 3 to 7 min for ISO VG46 hydraulic
fluids. Fluid D has a very low air release value of <1 min and is significantly lower
than other known commercial hydraulic fluids. This may have practical value for
hydraulic equipment that is designed in the future because original equipment
builders (OEM) continue to design systems with smaller sump reservoirs with
higher power densities. Fluids that do not release air fast enough can cause several
hydraulic performance problems including cavitation, sponginess, faster oxidation
of the fluid, and a reduction in the bulk modulus of the fluid. Fluid D may allow
GREAVES AND LARSON, DOI 10.1520/STP157320130103 137

FIG. 1 Air release times of commercially available hydraulic fluids that are ISO VG46.

equipment builders to reduce capital costs in designing smaller reservoirs. The


design of smaller reservoirs will mean that fluids are needed with even greater oxi-
dation stability than in the past. Fluid D shows excellent oxidation stability as
described above.
The manifold ignition test results are shown in Table 5. This test is usually per-
formed at 700 C. However, one can adjust the procedure so that a fluid can be
assessed for its flammability at different temperatures. The phosphate ester refer-
ence fluid did not burn at an ignition temperature of 700 C and showed superior
fire resistance than the ester, both PAGs, and the OSP fluid. The ester fluid had an
ignition temperature in the range of 373–425 C. Both PAG products and the OSP
fluid had an ignition temperature of 342–373 C. Interestingly, even though the
OSP fluid had a higher carbon to oxygen ratio in the polymer backbone, no differ-
entiation in ignition temperature performance was observed.
The ignitability index (RI), which is the primary measure of fire resistance from
the “stabilized flame heat release test” may also be used in a risk assessment to
determine a level of fire resistance appropriate to the operational situation. The test
results are shown in Table 5 and also in Fig. 2 and compared to a commercial phos-
phate ester. Higher values are preferred and indicate higher fire-resistance
performance.
In this test, a steady input of heat from a pilot burner flame is used to ignite a
defined spray of the fluid, but in this case the ignitability of the fluid is measured by
the amount of heat it releases. The temperature differences between the air entering
the test chamber and the exhaust gases are measured without fluid discharging
from the spray nozzle and also with fluid discharging from the spray nozzle. The
138 STP 1573 On Fire Resistant Fluids

FIG. 2 Ignitability values (RI) of fluids A to D and compared to a phosphate ester


reference fluid.

temperature difference is a measure of the heat released. An ignitability factor (RI)


is calculated on the basis of the relationship between the temperature differences,
i.e., the heat release (a) when only the burner is ignited, and (b) when the fluid
spray is ignited by the burner. The measurements are made when the flame pro-
duced by the burning spray has stabilized.
The highest RI value of 28 was for the phosphate ester. Water-based fluids such
as HFA, HFB, and HFC fluids have values that are much higher and are usually
>80. In contrast mineral oils have typical values <5. The ester product had a RI
value of only 5. The PAGs and OSP products showed higher values and in the range
9–11 suggesting marginally better performance. Data on the repeatability on this
test using these fluids is not available.
Based on these fire tests alone there is no differentiation between the OSP prod-
uct and the two PAG fluids, the latter of which have been used as fire-resistant flu-
ids for many years. This suggests OSPs could be potential building blocks of future
fire-resistant fluid formulations with some additional practical advantages as
described.

Conclusions
An assessment of the performance of non-phosphorus-based fire-resistant anhy-
drous hydraulic fluids that use either ester or polyalkylene glycol-based fluid com-
ponents has been conducted. Included in the assessment was a new oil-soluble
polyalkylene-glycol-based hydraulic fluid.
GREAVES AND LARSON, DOI 10.1520/STP157320130103 139

The ester-based fluid showed excellent viscosity index values and good low
temperature properties but its oxidation performance was inferior to the PAGs and
OSP fluids. All fluids showed excellent wear performance in a vane pump test
(ASTM D7043) but the ester showed viscosity deterioration after the test. Both
PAG fluids and the OSP fluid were stable. Examination of their air release proper-
ties showed the new OSP fluid technology to have significantly lower air release val-
ues. This feature alone may allow OEMs to design much smaller reservoirs and
possibly reduce capital costs. In an Eaton 35VQ25A vane pump test, the OSP fluid
demonstrated good wear performance and excellent stability.
The fire-resistance performance of each fluid using a hot manifold test and sta-
bilized flame heat release spray method showed inferior results when compared to a
reference phosphate ester. In the hot manifold test, the polyol ester fluid exhibited
higher values than the PAG and OSP fluids but lower ignitability factor values and
performance using the spray test. The OSP fluid did not show any significant differ-
ence on these tests versus the two PAG fluids that have been in commercial opera-
tion as fire-resistant fluids for several years. Because the OSP fluid has better
hydrocarbon oil compatibility and superior oxidation and air release properties,
and its performance in hydraulic equipment tests appears to be favorable, it may be
a better practical choice for future polyalkylene glycol-based fire-resistant hydraulic
fluids. Because OSP technology is new, further research work will determine suit-
ability for fire-resistance hydraulic applications.

ACKNOWLEDGMENTS
The writers thank many colleagues in The Dow Chemical Company who performed
the analytical and performance measurements in Freeport, TX, Midland, MI, and
Horgen, Switzerland.

References

[1] Webster, G. M. and Totten, G. E., “Fire Resistant Testing Procedures of Hydraulic Fluids,”
Handbook of Hydraulic Fluid Technology, G. E. Totten, Ed., Taylor and Francis, London,
2011, pp. 393–426.

[2] Snyder, C. E. and Gschwender, L. J., “Fire Resistant Hydraulic Fluids and Fire Resistance
Test Methods Used by the Air Force,” Fire Resistance of Industrial Fluids, ASTM STP
1284, G. E. Totten and J. Reichel, Eds., ASTM International, West Conshohocken, PA,
1996, pp. 72–77.

[3] Totten, G. E. and Bishop, R. J., “Historical Overview of the Development of Water-Glycol
Hydraulic Fluids,” Paper No. 952076, Society of Automotive Engineers, Warrendale, PA,
1995.

[4] Millett, W. H., “Fire Resistant Hydraulic Fluids,” Appl. Hydr., 1957, pp. 124–128.

[5] Sun, Y. and Totten, G. E., “Water Glycol Hydraulic Fluids,” Handbook of Hydraulic Fluid
Technology, G. E. Totten, Ed., Taylor and Francis, London, 2011, pp. 917–982.
140 STP 1573 On Fire Resistant Fluids

[6] Totten, G. E. and Webster, G. M., “High Performance Thickened Water Glycol Hydraulic
Fluids,” Proceedings of the 46th National Conference on Fluid Power, National Fluid
Power Association, Milwaukee, WI, 1994, pp. 185–194.

[7] ISO 6743-4: Lubricants, Industrial Oils and Related Products (Class L) Classification—
Part 4: Family H (Hydraulic Systems), International Standards Organization, Switzerland,
1999.

[8] Bishop, R. J. and Totten, G. E., “Maintenance and Analysis of Water Glycol Hydraulic Flu-
ids,” Mach. Lubr., Jan., 2003, pp. 24–31.

[9] ISO 12922: Lubricants, Industrial Oils and Related Products (Class L)—Family H (Hydrau-
lic Systems), Specifications for Categories HFAE, HFAS, HFB, HFC, HFDR and HFDU,
International Standards Organization, Switzerland, 1999.

[10] Marino, M. P., “Phosphate Esters,” Synthetic Lubricants and High Performance Functional
Fluids, R. L. Shubkin, Ed., Marcel Dekker, New York, 1992, pp. 67–100.

[11] Phillips, W. D. and Sutton, D. I., “Improved Maintenance and Life Extension of Phosphate
Esters Using Ion Exchange Treatment,” J. Synth. Lub., Vol. 13(3), 1996, pp. 225–261.

[12] Anglin, J., “Fire Resistant Hydraulic Fluids for Aluminum Processing,” STLE National Con-
ference, Cleveland, OH, May 18–22, 2008.

[13] Phillips, W. D., “A Comparison of Fire Resistant Hydraulic Fluids for Hazardous Applica-
tions, Part 2,” J. Synth. Lub., Vol. 14(4), 1998, pp. 307–330.

[14] Greaves, M. R. and Knoell, J. C., “A Comparison of the Performance of Environmentally


Friendly Anhydrous Fire Resistant Hydraulic Fluids,” J. ASTM. Int., Vol. 6(10), 2009, pp.
1–10.

[15] Factory Mutual Research Corp., “Flammability Classification of Industrial Fluids, Class
6930,” Factory Mutual Corp., Approved Standard, 2002.

[16] “Requirements and Tests Applicable to Fire Resistant Hydraulic Fluids Used for Power
Transmission and Control,” The 7th Luxembourg Report, Doc. No. 4746/10/91, Safety
and Health Commission for the Mining and Other Extractive Industries, L-2920m, Lux-
embourg Commission of the European Economic Communities, DG\E\4, 1994.

[17] Greaves, M. R., van Voorst, R., Meertens, R., Zaugg-Hoozemans, E., and Khelidj, N.,
“Performance Properties of Oil-Soluble Synthetic Polyalkylene Glycols,” Lubr. Sci., Vol.
24(6), 2012, pp. 251–262.

[18] Greaves, M. R., “Oil Soluble Synthetic Polyalkylene Glycols,” Lube Magazine, Aug 2011,
pp. 21–24.

[19] ASTM D445-12: Standard Test Method for Kinematic Viscosity of Transparent and Opa-
que Liquids, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2013.

[20] ASTM D2270-10e1: Standard Practice for Calculating Viscosity Index from Viscosity at
40 and 100 C, Annual Book of ASTM Standards, ASTM International, West Consho-
hocken, PA, 2013.

[21] ASTM D97-12: Standard Test Method for Pour Point of Petroleum Products, Annual Book
of ASTM Standards, ASTM International, West Conshohocken, PA, 2013.
GREAVES AND LARSON, DOI 10.1520/STP157320130103 141

[22] ASTM D92-12b: Standard Test Method for Flash and Fire Points by Cleveland Open Cup
Tester, Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2013.

[23] ASTM D7042-12a: Standard Test Method for Dynamic Viscosity and Density of Liquids
by Stabinger Viscometer, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2013.

[24] ASTM D664-11a: Standard Test Method for Acid Number of Petroleum Products by
Potentiometric Titration, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2013.

[25] ASTM E1719: Standard Method for Vapor Pressure of Liquids by Ebulliometry, Annual
Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2013.

[26] Matlock, P. L. and Clinton, N. A., Polyalkylene Glycols, R. L. Shubkin, Ed., Marcel Dekker,
New York, 1993, pp. 101–123.

[27] Greaves, M. R., “Polyalkylene Glycols,” Synthetics, Mineral Oils and Bio-Based Lubricants,
L. R. Rudnick, Ed., CRC, Boca Raton, FL, pp. 123–148.

[28] ASTM D7043-12: Standard Test Method for Indicating Wear Characteristics of Petroleum
and Non-Petroleum Hydraulic Fluids in a Constant Vane Pump, Annual Book of ASTM
Standards, ASTM International, West Conshohocken, PA, 2013.

[29] ASTM D6973-05: Standard Test Method for Indicating Wear Characteristics of Petro-
leum Hydraulic Fluids in a High Pressure Constant Volume Vane Pump, Annual Book of
ASTM Standards, ASTM International, West Conshohocken, PA, 2013.

[30] ASTM D665-12: Standard Test Method for Rust-Preventing Characteristics of Inhibited
Mineral Oil in the Presence of Water, Annual Book of ASTM Standards, ASTM Interna-
tional, West Conshohocken, PA, 2013.

[31] ISO-2160-98, “Petroleum Products – Corrosiveness to Copper – Copper Strip Test,”


1998.

[32] ASTM D2272: Standard Test Method for Oxidation Stability of Steam Turbine Oils by
Rotating Pressure Vessel, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2013.

[33] ASTM D2893-04: Standard Test Method for Oxidation Characteristics of Extreme-
Pressure Lubrication Oils, Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2013.

[34] ASTM D3427-04: Standard Test Method for Air Release Properties of Petroleum
Oils, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2013.

[35] ASTM D1401-12: Standard Test Method for Water Separability of Petroleum Oils and Syn-
thetic Fluids, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2013.

[36] ASTM D4172-94: Standard Test Method for Wear Preventive Characteristics of Lubricat-
ing Fluid (Four-Ball Method), Annual Book of ASTM Standards, ASTM International,
West Conshohocken, PA, 2013.
142 STP 1573 On Fire Resistant Fluids

[37] ISO 15029: Petroleum and Related Products—Determination of the Spray Ignition Char-
acteristics of Fire Resistant Fluids—Part 2: Stabilized Flame Heat Release Method, Inter-
national Standards Organization, Geneva, Switzerland, 2012.

[38] ISO 20823: Petroleum and Related Products—Determinations of the Flammability Char-
acteristics of Fluids in Contact with Hot Surfaces—Manifold Ignition Test, International
Standards Organization, Geneva, Switzerland, 2003.

[39] Greaves, M. R. and Khemchandani, G., “Novel Polyalkylene Glycol-Based Hydraulic Flu-
ids,” Proceedings of the Iron and Steel Technology Conference, Dec 2010, pp. 66–71.
FIRE RESISTANT FLUIDS 143

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130190

Kevin P. Kovanda1 and Mark Latunski2

Polyalkylene Glycol Hydraulic


Fluids, 20 Years of Fire Resistance
Reference
Kovanda, Kevin P. and Latunski, Mark, “Polyalkylene Glycol Hydraulic Fluids, 20 Years of Fire
Resistance,” Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 143–154, doi:10.1520/
STP157320130190, ASTM International, West Conshohocken, PA 2014.3

ABSTRACT
Fluid Power is the preferred choice by industry to perform work due to its
smaller size, higher energy efficiency and ease of adjustment. All of these
advantages are lost if an incident occurs in which the hydraulic fluid, under
pressure, is sprayed in the presence of an ignition source resulting in a fire, harm
to workers or the environment.

Keywords
polyalkylene glycols, fire resistance, properties of polyglycols

Introduction
Fluid power is the preferred choice by industry to perform work because of its smaller
size, higher energy efficiency, and ease of adjustment. All of these advantages are lost if
an incident occurs in which the hydraulic fluid, under pressure, is sprayed in the pres-
ence of an ignition source resulting in a fire or harm to workers or the environment.
Petroleum oils are the most commonly used hydraulic fluid, but some applica-
tions demand a higher degree of fire resistance. In these situations, fire-resistant flu-
ids are used. However, fire-resistant fluids are not completely inflammable, most
organic materials will burn under the right conditions. From the viewpoint of insur-
ance underwriters, labor organizations, government regulations, and the fluid

Manuscript received December 17, 2013; accepted for publication June 10, 2014; published online July 18,
2014.
1
President, American Chemical Technologies, Inc., Fowlerville, MI 48836, United States of America, e-
mail: kkovanda@americanchemtech.com
2
Laboratory Manager, American Chemical Technologies, Inc., Fowlerville, MI 48836, United States of America.
3
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
144 STP 1573 On Fire Resistant Fluids

power industry itself, it is becoming increasingly critical to be able to determine


appropriately the relative fire resistance provided by the use of a particular fluid in
a specific industrial process.
Polyalkylene glycols (PAGs) have been meeting these requirements for more
than 20 years.

Fire Triangle
The fire triangle [1] or combustion triangle is a simple model for understanding the
necessary ingredients for most fires. Figure 1 shows the fire triangle with its three ele-
ments needed for ignition, with the corresponding influence that a hydraulic fluid
basestock, polyalkylene glycol (PAG), polyol ester (ester), and phosphate ester (PE),
has on suppressing that ignition. ISO 6743-4 [2] classifies fire-resistant hydraulic
fluids into water-containing (HFA, HFB, and HFC) and non-water-containing cate-
gories (HFD). The non-water-containing category is further divided into HFDR,
synthetic fluids containing no water and phosphate esters, and HFDU, synthetic
fluids containing no water and other components, which could include polyol esters
and polyalkylene glycols. Table 1 illustrates how the non-aqueous fire-resistant hy-
draulic fluids, categories HFDR and HFDU, chemically fall into these elements.
Fire-resistant testing and approval is governed by FM Global, and in the Euro-
pean Union via ISO-12922 [3]. Both testing bodies and industry in general agree
that there is more risk associated with combustion than just fluid flammability. All
would agree that fires generate two types of fire products in an uncontrolled fash-
ion: heat and chemical compounds [4]. The most often ignored issue is the poten-
tial toxicity of fluid combustion by-products that may be formed [5]. Fire resistance
is one aspect, but it was the toxicity of possible pyrolysis products from halogenated

FIG. 1 Fire triangle.


KOVANDA AND LATUNSKI, DOI 10.1520/STP157320130190 145

TABLE 1 Non-aqueous fire-resistant hydraulic fluids.

Molecular Weight Fire Point Heat of Combustion

Ester Good Good Fair


PAG Good Fair Good
PE Fair Good Good

hydrocarbons and the severe effects they may have on humans and the environ-
ment that eventually prohibited their use [6]. Otherwise, polychlorinated biphenyls
(PCBs) would still be in use today. It was their potential biological and toxicological
hazard that was their demise. That is why both testing bodies state: FM—“this
standard is intended only to evaluate a fluid’s flammability under stated conditions.
Environmental considerations, toxicity and the suitability for the end use of the
product have not been evaluated”; and ISO—“This international standard does not
purport to address all of the safety problems associated with the use of fire-resistant
fluids. It is the responsibility of users to establish appropriate safety and health
practices and to determine the applicability of regulatory limitations prior to use.
The fluid shall not present any significant hazard to health when correctly used in
hydraulic equipment, observing the handling recommendations of the supplier.”
Potential fire resistance and environmental and toxicological properties of flu-
ids are composition dependent. They will be influenced by the chemical structure of
the fluid and the geometry or configuration of the fluid fire. Chemical compounds
generated by fluid fires are further distributed into substances associated with com-
plete combustion (carbon dioxide and water vapor) and substances resulting from
incomplete combustion (carbon monoxide, particulates, and gaseous degradation
products—oxides of nitrogen, phosphorus, etc.) [7]. Figure 2 represents the laby-
rinth that an end user must go through in their decision-making process to choose
a fire-resistant hydraulic fluid.

PAGs—The Newest Member


of the Fire-Resistant Family [8]
PAGs are unique in that they represent a class of lubricants that not only pass strin-
gent environmental tests, they also provide the end user with unmatched long-term
performance and fire resistance. For over 50 years, PAGs have been solving prob-
lems that hydrocarbon lubricants cannot.
PAGs can be chemically designed to meet an extensive range of performance
needs. No other base oil chemistry has such versatility. Generally speaking, a PAG
(also known as a polyglycol, a polyol, and a polyether) is prepared by reacting an
initiator with one or more alkylene oxides under alkaline conditions and elevated
temperatures. Within this reaction, there are four variables: the initiator, the oxides,
the way the oxides are reacted on by the initiator (i.e., random or block additions),
146 STP 1573 On Fire Resistant Fluids

FIG. 2 Decision-making process when choosing fire-resistant hydraulic fluid.

and the molecule weight. This gives PAG chemistry its versatility, which makes it
possible to create tailor-made products that meet desired requirements.
The initiator gives the PAG its “chemical functionality,” and influences the
physical properties of the resulting product, for example, by imparting hydropho-
bicity to the polymer. Most initiators are alcohols, such as butanol, mono-
propylene glycol, or glycerine. Oxides react with the labile hydrogens of the
initiators to produce linear or branched structures depending on the functionality
of the initiator. The choice of oxides will influence the hydrophilic/hydrophobic
character of the resulting PAG. The most commonly used oxides include ethylene
oxide, propylene oxide, and butylene oxide. A high content of ethylene oxide will
normally result in fully water-soluble products, whereas a high content of butylene
oxide will provide oil solubility. By combining the oxides, the solubility behavior of
the products can be adjusted. The molecular weight of the PAG will affect the vis-
cosity of the product. With PAG chemistry, almost the complete ISO VG viscosity
range can be covered.

Properties of Polyglycols
VISCOSITY
With PAG chemistry, a broad range of the ISO VG viscosity classes can be covered
(Fig. 3). Typical standard grades range from ISO 10 to ISO 1000. As indicated
above, the viscosity of a polyglycol depends mainly on the molecular weight; the
viscosity will, in general, increase with increasing molecular weight.
KOVANDA AND LATUNSKI, DOI 10.1520/STP157320130190 147

FIG. 3 PAG chemistry.

VISCOSITY INDEX
When compared with other base fluids, PAGs show, in general, very high viscosity
indexes. VI values up to 400 can be reached (Fig. 4).

SOLUBILITY
As mentioned above, it is possible to create polyglycols with solubility properties
ranging from complete water solubility to complete oil solubility.

LUBRICITY
Polyglycols show excellent lubrication properties because of the high affinity of the
oxygen atoms in the polymer with the metal surface. This feature also provides
mild extreme pressure properties.

THERMAL STABILITY
Polyglycols demonstrate a very good response to antioxidants, and can therefore be
formulated for high-temperature applications (i.e., up to about 250 C). Other
148 STP 1573 On Fire Resistant Fluids

FIG. 4 Very high viscosity indexes of PAGs.

basestocks, such as vegetable oils, some synthetic esters, mineral oils, and PAOs, do
not offer good enough thermal stability in this temperature range.

LOW RESIDUE
Polyglycols show superior deposit control characteristics over all other base oil
classes. Oxidation of polyglycols results in a partial breakdown of the product into
volatile components and polar compounds soluble in the base fluid, resulting in low
formation of residue. In the case of most other basestocks, oxidation normally leads
to polymerization, which produces high molecular weight polar by-products that
are insoluble in their parent base oil leading to significant residues, which can be an
important disadvantage in certain applications (Fig. 5).

LOW POUR POINTS


Pour points as low as 45 C can be obtained with polyglycol chemistry. This
makes polyglycols suitable for low-temperature applications. Low pour points com-
bined with their high viscosity indexes make polyglycol lubricants an excellent
choice for applications requiring one product for all seasons.

THERMAL CONDUCTIVITY
In general, PAGs offer better thermal conductivity than mineral oils. This means
that better cooling characteristics can be achieved with PAGs, and, as a result, a
smaller heat exchanger can be used.

HYDROLYTIC STABILITY
Polyglycols do not hydrolyze, which can be seen as a major advantage over natural
esters (vegetable oils), synthetic esters, and phosphate esters (Fig. 6). With esters,
KOVANDA AND LATUNSKI, DOI 10.1520/STP157320130190 149

FIG. 5 Results of oxidation of polyglycols.

hydrolysis leads to acid formation, which will increase the lubricant’s corrosion
potential and accelerate further fluid degradation. In the case of phosphate esters,
this hydrolysis is autocatalytic, meaning that the acid formed will now act as a cata-
lyst to form more acid [9]. In many applications, contamination with water cannot
be completely avoided. This is especially true for equipment operating with envi-
ronmental concerns should there be a leak. Most of this equipment operates out-
side, exposed to the elements where rain and humidity ingress into the lubricant
system; and this is how the hydrolysis process starts. For polyglycols, however, this
is not a problem. A further unique feature is their ability to absorb water that con-
taminates the lubricant through ingress. When this occurs in the equipment, the
PAG can act as a polymeric sponge, which renders the water inert.

ELASTOMER/METAL COMPATIBILITY
Care should be taken in the choice of elastomers with all base fluids. A broad range
of commercially available elastomers, such as NBR (acryonitrite/butadiene rubber)
and FPM (fluoro-rubber), can be used with polyglycols. The same is true for ferrous
and non-ferrous metals.

FIRE RESISTANCE
As was shown in the fire triangle, molecular weight plays a role in fire resistance.
PAGs, depending on their molecular weight, can be formulated to pass FM Global
6930 [10] and ISO-12922 [3]. PAGs received their first fire-resistance approval in 1989.
Table 2 highlights typical fluid properties for a water-insoluble PAG that is fire-
resistance approved. Table 3 lists the current industries being served by PAG fire-
resistant hydraulic fluids.
150 STP 1573 On Fire Resistant Fluids

FIG. 6 Hydrolytic stability (ASTM D943-04a(2010)e1 [11]).

Slag Industry—First Fire-Resistant Application


for PAGs, 1991
The first application for a PAG, fire-resistant hydraulic fluid was in a 992 C wheel
loader in 1991 (Fig. 7). The machine had previously run on a synthetic ester fluid

TABLE 2 Typical fluid properties for a water-insoluble PAG.

Properties ISO VG 46 ISO VG 68


Viscosity at 40 C (cSt) 50 68
Viscosity at 100 C (cSt) 9.6 12.6
Viscosity index 180 185
Density at 15 C 0.985 0.986
Pour point,  C 48 45
Copper strip corrosion 1a, shiny 1a, shiny
Rust test, synthetic sea water Pass Pass
TOST (with water) (2000 h) Pass Pass
Foam test (sequences I, II, and Pass Pass
III)
Air release (min) 8.1
Four ball wear, scar (mm) 0.35 0.35
FZG (stages passed) 12 12
Vane pump test Pass Pass
Brugger value >40 N/mm2 >40 N/mm2
Readily biodegradable Yes Yes
FM approved (fire resistant) Yes Yes
KOVANDA AND LATUNSKI, DOI 10.1520/STP157320130190 151

TABLE 3 Current industries served by PAG fire-resistant hydraulic fluids.

Mobile equipment—slag industry


Steel
Aluminum
Casting
Tunnel boring
Glass
Power generation
Fossil
Hydro

and pump life averaged 2000 h. The PAG fluid was put into this machine, which
had “used” pumps, and they ran for 9000 h. Life on new pumps has consistently
averaged 10 000–12 000 h. This was a significant improvement considering the
machine operates at 225 bar (3250 psi) with a very high flow rate of 1410 l/min
(310 gpm) with a small reservoir volume of 636 l (140 gal).
Figure 8 shows a Caterpillar 973C with the steel mill option. A water-insoluble
PAG was chosen as the fluid for initial-fill based on its performance and versatility.
An ISO-46 grade fluid is used in the hydrostatic drives, implement hydraulics, and
splitter gearbox. Pressure relief settings are 450 bar (6500 psi) for the hydrostatic
drive system to provide high breakout force.

FIG. 7 992C wheel loader.


152 STP 1573 On Fire Resistant Fluids

FIG. 8 Caterpillar 973C with the steel mill option.

Steel Industry—17-Year History


PAGs, because of their hydrolytic stability, low residue formation, and excellent
lubrication, have established a 17-year track record of impeccable service in the steel
industry, most noticeably in hydraulic automatic gauge control systems (HAGC).
PAGs are now the fluid of choice in these applications because of the high operating
pressures (5000 psi) and tight tolerance valves (servos) used for the instantaneous

FIG. 9 An insignificant increase in acid number after a water leak developed.


KOVANDA AND LATUNSKI, DOI 10.1520/STP157320130190 153

FIG. 10 Seasonal fluctuation of water content (ASTM D-664-11a [12]).

control of the “squeeze” applied to the steel strip as it is passing through the finish-
ing stands. Pump life of 85 000 h has been obtained with the use of PAGs, whereas
historical life of only 4000 h was obtained with synthetic esters. Figures 9 and 10
highlight the hydrolytic stability of this water-insoluble PAG as acid number versus
water content is graphed since the inception of the fluid use. Note the insignificant
increase in acid number even after a water leak developed (6 %—60 000 ppm) in
Fig. 9. This water ingression was treated on-line with a vacuum dehydrator and the
line was never shutdown. No servo valve failures occurred because of this leak ei-
ther. Figure 10 shows how the water content can fluctuate seasonally and yet there is
no corresponding increase in acid number as would be experienced with an ester-
based fluid.

Conclusion
There is no perfect fire-resistant hydraulic fluid. Each chemical class has its strong
points and weak points. In evaluating the classes of fire-resistant hydraulic fluids,
HFDU-PG (PAGs) represents the overall most user-friendly technology available.

References

[1] “Fire Triangle,” en.wikipedia.org/wiki/Fire_Triangle (Last accessed 25 June 2006).

[2] ISO 6743-4: Lubricants, Industrial Oils and Related Products (Class L) – Classification –
Part 4: Family H (Hydraulic systems), ISO, Geneva, Switzerland, 1999.
154 STP 1573 On Fire Resistant Fluids

[3] ISO-12922: Lubricants, Industrial Oils and Related Products (Class L) – Family H (Hydrau-
lic Systems) – Specifications for Hydraulic Fluids in Categories HFAE, HFAS, HFB, HFC,
HFDR and HFDU, ISO, Geneva, Switzerland, 2012.

[4] Newman, J. S., “Combustion Fire Chemistry of Industrial Fluids,” Fire Resistance of
Industrial Fluids, ASTM STP 1284, G. E. Totten and J. Reichel, Eds., ASTM International,
West Conshohocken, PA, 1996.

[5] Totten, G. E., “Overview,” Fire Resistance of Industrial Fluids, ASTM STP 1284, G. E. Totten
and J. Reichel, Eds., ASTM International, West Conshohocken, PA, 1996.

[6] Reichel, J., “Standardization Activities for Testing of Fire Resistant Fluids,” Fire Resist-
ance of Industrial Fluids, ASTM STP 1284, G. E. Totten and J. Reichel, Eds., ASTM Interna-
tional, West Conshohocken, PA, 1996.

[7] Phillips, W. D., “Fire Resistance Test for Fluids and Lubricants—Their Limitations and
Misapplication,” Fire Resistance of Industrial Fluids, ASTM STP 1284, G. E. Totten and
J. Reichel, Eds., ASTM International, West Conshohocken, PA, 1996.

[8] van Voorst, R. and Alam, F., “Polyglycols as Base Fluids for Environmentally Friendly
Lubricants,” 11th International Tribology Colloquium “Industrial and Automotive
Lubrication,” Jan 13–15, 1998, Technische Akademie Esslingen, Ostfildern, Germany.

[9] Phillips, W. D., “Phosphate Ester Hydraulic Fluids,” Handbook of Hydraulic Fluid Technol-
ogy, G. E. Totten, Ed., Union Carbide Corporation, Tarrytown, NY, 2000.

[10] FM Approval Standard: Flammability Classification of Industrial Fluids, FM Global Group,


Norwood, MA, 2009.

[11] ASTM D943-04a(2010)e1: Standard Test Method for Oxidation Characteristics of Inhib-
ited Mineral Oils, Annual Book of ASTM Standards, West Conshohocken, PA, 2010.

[12] ASTM D664-11a: Standard Test Method for Acid Number of Petroleum Products by
Potentiometric Titration, Annual Book of ASTM Standards, West Conshohocken, PA,
2011.
FIRE RESISTANT FLUIDS 155

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320130101

S. Rea1 and D. Barker2

Performance Comparison of
Non-Aqueous Fire-Resistant
Hydraulic Fluids
Reference
Rea, S. and Barker, D., “Performance Comparison of Non-Aqueous Fire-Resistant Hydraulic
Fluids,” Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 155–180, doi:10.1520/
STP157320130101, ASTM International, West Conshohocken, PA 2014.3

ABSTRACT
The performance of a hydraulic fluid has a significant impact on the real-world
behavior of the fluid in critical applications, such as electro-hydraulic control
systems in nuclear facilities. This paper examines some of the critical
performance parameters that determine the suitability of a fluid in these
demanding environments. A representative selection of non-aqueous fluid types
were chosen based on their individual chemistry and that are in use within
industry today. These fluids were then subjected to a series of performance tests
based on measuring their flammability, their lubricity, and their ability to function
as an effective hydraulic fluid under varying conditions. The results are presented
and explored in detail, with recommendations as to their suitability for use within
high-risk application areas.

Keywords
fire resistant non-aqueous hydraulic fluids, triaryl phosphate ester, polyol ester,
polyalkylene glycol, PAO/diester, mineral oil, anti-wear, lubricity, falex pin-and-
vee, four-ball, Cameron–Plint, fire resistance tests

Manuscript received June 9, 2013; accepted for publication September 26, 2013; published online December
27, 2013.
1
Anderol Specialty Lubricants, East Hanover, NJ 07936 (Corresponding author).
2
Chemtura Europe, Langley, Surrey, SL3 6EH, UK.
3
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
156 STP 1573 On Fire Resistant Fluids

Introduction
Hydraulic fluids are used in a broad range of applications and can be found in steel
mills, paper mills, coal, and metal ore mines, metal ore processing plants, construc-
tion (mobile) equipment, power plant control systems for both steam and gas tur-
bines, nuclear power plants, commercial aviation equipment, and in military
vehicles (aviation and non-aviation).
Fire resistance of hydraulic fluids remains a critical fluid property in many
industries where fluid exposure to high temperature and an ignition source could
result in a catastrophic fire, e.g., in an underground coal mine or in a nuclear
power plant. Fluid leaks can lead to spraying or pooling of liquid. Fire resistance
must take into account not only the ignitability of a fluid in these situations, but
also the combustion properties of a fluid once ignited. The combustion properties
are directly related to the potential of a fire to spread away from the original igni-
tion site.
Demands on the hydraulic fluid in all these applications continue to increase.
Sump sizes are shrinking, pump speeds and fluid operating pressures and tempera-
tures are increasing. As a result, fluid properties are being increasingly stressed in
modern hydraulic systems.
Besides fire resistance, there are several other key properties to consider when
selecting a hydraulic fluid for a specific application. These include anti-wear/lubric-
ity, environmental toxicity and biodegradability, air and water separability, hydro-
lytic, shear, oxidative and thermal stability, viscosity grade, cleanliness (particle
count), and, finally, cost. The choice of hydraulic fluid needs to be tailored to the
application. For example, hydraulic fluids used as electro-hydraulic control (EHC)
fluids in power plants must meet particularly stringent air and water separability as
well as hydrolytic stability requirements.
The different general types of fire resistant hydraulic fluids available to an end
user are classified as per ISO 6743-4 [1]. These comprise both aqueous and non-
aqueous fluids. The aqueous fluids include: HFA-E (oil-in-water emulsions), HFA-
S (chemical solutions in water), HFB (water-in-oil emulsions), and HFC (water
polymer solutions).
This study focuses on the non-aqueous fluid classes as defined by ISO 6743-4
[1]: HFDR (phosphate esters) and HFDU (all others, including carboxylic acid
esters, polyalkylene glycols, polyalphaolefin diester blends, and vegetable oils). A set
of performance data was compared for eight commercially available non-aqueous
hydraulic fluids. Testing included physicals and chemicals, air and water separabil-
ity, hydrolytic stability, anti-wear/lubricity, and fire resistance.
Fire resistance in this study was measured by running several different special-
ized bench tests as specified by ISO 12922 [2]. These include: ISO 20823 (“hot man-
ifold” test [3]), ISO 14935 (“wick flame persistence” test [4]), and ISO 15029-1 and
ISO 15029-2 (“spray flame” tests [5,6]).
REA AND BARKER, DOI:10.1520/STP157320130101 157

Anti-wear/lubricity properties are normally measured using expensive hydrau-


lic pump tests (e.g., the Vickers rotary vane pump test and the Parker–Denison
hybrid vane/axial piston pump test). Sometimes less expensive bench tests are used
as screeners to predict performance. Several different anti-wear/lubricity bench tests
were used in this study to compare the fluids, including: Falex Pin-and-Vee, Four-
Ball Wear, and Cameron–Plint wear and friction tests.

COMPARATIVE STUDY OF EIGHT COMMERCIALLY AVAILABLE NON-


AQUEOUS HYDRAULIC FLUIDS
Eight commercially available non-aqueous hydraulic fluids were chosen for this
study. All are the same viscosity grade (ISO VG 46). All are fully formulated fluids,
comprising a base stock and additives. Five different base stocks were examined: tri-
aryl phosphate ester, polyol ester, polyalkylene glycol, polyalphaolefin and diester
blend, and mineral oil.
For purposes of identification in this paper, the eight fluids are coded as
follows:
• PHE1 Triaryl phosphate ester (xylenated)
• PHE2 Triaryl phosphate ester (butylated phenyl)
• PHE3 Triaryl phosphate ester (butylated phenyl)
• POE1 Polyol ester (TMP oleate)
• POE2 Polyol ester (TMP oleate)
• PAG Polyalkylene glycol (polypropylene glycol, water insoluble)
• PAOE Polyalphaolefin (PAO) and diester (phthalate ester)
• MO Mineral oil (Group I)

PHE1, PHE2, PHE3, and PAG claim to be fire resistant and are marketed as
EHC fluids.
PAG also claims to be a biodegradable fluid.
POE1 and POE2 claim to be fire resistant and biodegradable fluids.
PAOE and MO are marketed as general purpose fluids and do not claim fire
resistance.
Chemical structures are shown here for the ester and polypropylene glycol base
stocks:
Xylyl phosphate ester (PHE1 base stock)
158 STP 1573 On Fire Resistant Fluids

Tert-butyl phenyl phosphate ester (PHE2 and PHE3 base stock)

TMP (trimethylol propane) oleate ester (POE1 and POE2 base stock)

Polypropylene glycol (PAG base stock)

Phthalate ester (ester base stock of PAOE)


REA AND BARKER, DOI:10.1520/STP157320130101 159

TABLE 1 Physical and chemical test data.

IEC 60 247
Coded ISO ASTM D2270 ISO 6619 ISO 760 IP 510.04 ISO 4406 DC Current
Fluid 3104 KV, Viscosity TAN, mg Water, Chlorine, Particle Resistivity,
Name mm2/s Index KOH/g mass % mppm Count MX-m

PHE1 44.2 <0 0.04 0.03 <2 19/2013/8 346


PHE2 42.7 <0 0.02 0.05 <2 15/13/8 354
PHE3 43.5 <0 0.03 0.03 <2 18/16/11 243
POE1 48.5 195 1.22 0.02 4 14/12/8 >1000
POE2 47.4 190 0.78 0.01 <2 16/13/10 >1000
PAG 49.5 180 0.19 0.08 <2 17/14/8 40
PAOE 43.7 131 0.07 0.01 <2 17/14/10 0.1
MO 45.3 97 0.11 <0.01 <2 17/13/9 0.5

Physical and Chemical Data


Table 1 shows some physical and chemical characterization data measured for the
eight fluids: kinematic viscosity, viscosity index, total acid number (TAN), water
content, chlorine content, and particle count.
All eight fluids are ISO VG 46 (i.e., 46 mm2/s kinematic viscosity at 40 C, 610
%). However, they vary widely in viscosity index (VI), ranging from less than zero
for the phosphate esters to nearly 200 for the polyol esters and the polyalkylene
glycol.
The phosphate esters have the lowest TAN values, while the polyol esters have
the highest TAN values. ISO 12922 does not specify any limits for TAN (report
only).
The water content for the polyalkylene glycol is the highest at 800 ppm. Water
contents range from 100 to 500 ppm for the other fluids. ISO 12922 specifies a max-
imum 1000 ppm water content.
The chlorine content for all eight fluids are very low, just above or below the
detection limit.
The particle counts for these fresh fluids are all relatively low (as expected). Hy-
draulic systems can typically include a filtration unit to maintain cleanliness of in-
service fluids.
Electrical resistivity is an important property for electrohydraulic control
(EHC) fluids. Low resistivity fluids can be characterized by electric currents which
cause wear damage on critical, tight tolerance surfaces inside servo valves [7]. The
minimum DC current resistivity for unused triaryl phosphate ester fluids in EHC
service as specified by ISO 10050 [8] is 50 MX-m. The three PHE fluids all easily
meet this requirement. The PAG fluid, which also claims to be an EHC fluid, does
not meet this minimum resistivity requirement.
160 STP 1573 On Fire Resistant Fluids

TABLE 2 Air and water separability test data.

ASTM ISO 6247 ISO 6247 ISO 6247


Coded ISO 9120 D1401 Water Foam Seq I Foam Seq II Foam Seq III
Fluid Air Release Demulse Tendency/Stability, Tendency/Stability, Tendency/Stability,
Name Time, Minutes Time, Minutes mls/mls mls/mls mls/mls

PHE1 1.0 5 0/0 0/0 40/0


PHE2 1.0 5 20/0 20/0 20/0
PHE3 3.8 5 210/0 20/0 270/0
POE1 5.8 15 50/0 540/0 10/0
POE2 0.6 20 0/0 0/0 0/0
PAG 2.4 >30 530/0 40/0 470/0
PAOE 1.1 10 80/0 20/0 20/0
MO 4.1 20 100/0 20/0 200/0

Air and Water Separability Data


Table 2 shows some air and water separability data for the eight fluids: air release
time, water demulse time, and foam tendency/stability.
Air release times are all under the maximum 15 min required by ISO 12922 for
ISO VG 46 HFDR and HFDU hydraulic fluids. In addition, all eight fluids also give
air release times under the maximum 6 min required by ISO 10050 [8] for ISO VG
46 triaryl phosphate ester turbine control fluids.
Water demulse times for the three phosphate esters are 5 min and easily meet
the maximum 15 min allowed by ISO 10050. The POE1 and PAOE fluids also gave
low water demulse times (15 and 10 min, respectively). The PAG fluid gave the lon-
gest demulse time (>30 min). Even though this fluid is a “water insoluble” polyalky-
lene glycol, the water solubility is high enough to form a stable emulsion in this test.
The foam stability results for all eight fluids is excellent (0 mls). However, the
foam tendency results are varied. The PAG fluid does not meet the 300/10, 300/10,
300/10 maximum foam requirements of ISO 12922, with 530 ml and 470 ml foam-
ing tendency in Seq I and Seq III, respectively. All the other fluids meet the ISO
12922 foam requirements, except for the POE1 fluid, which exhibited a very high
and repeatable Seq II foam tendency result. Of the three phosphate esters, only flu-
ids PHE1 and PHE2 meet the 150/0, 30/0, 150/0 foam requirements of ISO 10050.

Hydrolytic Stability Data


Table 3 shows the ASTM D2619-09 [9] hydrolytic stability test results for the eight
fluids: copper weight change, copper panel appearance, change in acid number of
the oil, acidity of the aqueous layer, % kinematic viscosity change, and % insolubles.
Test repeatability as given in the standard for three of the test responses: 60.3 mg/
cm2 copper corrosion, 60.8 mg KOH/g acid number change of the oil,
and 6 0.8 mg KOH/g total acidity of water layer.
REA AND BARKER, DOI:10.1520/STP157320130101 161

TABLE 3 Hydrolytic stability data (ASTM D2619-09).

Change Acidity of
Coded Copper Weight Copper in D974 Water Kinematic
Fluid Change, Panel TAN, Layer, Viscosity %
Name mg/cm2 Rating mg KOH/g mg KOH/g % Change Insolubles

PHE1 –0.125 1B 0.00 0.54 –4.56 0.007


PHE2 –0.017 1B 0.00 1.15 –8.09 0.017
PHE3 –0.125 1B 0.02 3.19 –6.77 0.006
POE1 –0.025 1B 0.86 9.24 –1.00 0.024
POE2 –0.050 2C 0.35 4.89 þ0.42 0.004
PAG –0.008 1B 0.05 6.11 –7.90 0.027
PAOE 0.000 1B 0.00 1.29 –0.18 0.011
MO –0.167 1B –0.02 1.70 þ0.20 0.003

Overall, the PAOE fluid performed the best across all test responses. POE1
gave the highest change in acid number and highest water layer acidity. This sug-
gests a greater susceptibility to hydrolysis compared to the other esters (even com-
pared to POE2 which is also a TMP oleate).
Note that the change in acid number values for all the oils are nearly within the
range of repeatability and so the differences are not statistically significant at the 95
% confidence level.
Copper corrosion rates were all acceptable and equivalent within the repeatabil-
ity range. The POE1 and PAG fluids gave the highest level of insolubles.
ISO 12922 does not set any limits for hydrolytic stability (report only require-
ment); however, ISO 10050 specifies a maximum increase in acid number of 0.5 mg
KOH/g for triaryl phosphates in turbine control fluid service. All three phosphate
ester fluids meet this requirement easily.

SUMMARY OF PHYSICAL AND CHEMICAL, AIR AND WATER SEPARABILITY,


AND HYDROLYTIC STABILITY
All eight fluids are ISO VG 46 with low water and chlorine contents. Viscosity
index and total acid number values vary, with the triaryl phosphate esters having
the lowest VI and TAN values. All the fluids are relatively clean with low particle
counts.
Air and water separability are key performance attributes for hydraulic fluids,
especially for EHC fluids, where sensitivity to air and water release is enhanced by
the need to maintain extremely precise servo valve control. Poor air or water release
will lead to poor valve control and other issues such as oxidation deposits and poor
lubricity.
Air release times for all eight fluids are relatively low.
The triaryl phosphate esters have the lowest water demulse times. The PAG
fluid formed a stable emulsion in this test, indicating that it would not provide
adequate water shedding performance.
162 STP 1573 On Fire Resistant Fluids

Foam stability for all eight fluids is excellent; however, foam tendency for the
PAG and POE1 fluids are relatively high. For EHC fluid service, PHE1 and PHE2
meet the foam tendency requirements of ISO 10050, while PHE3 does not meet the
requirements.
Hydrolytic stability is another key performance attribute for hydraulic fluids.
Poor hydrolytic stability can lead to formation of corrosive acidic species and
deposits.
The PAOE fluid gave the best overall hydrolytic stability results. The POE1
fluid gave the highest change in acid number and highest water layer acidity. This
suggests that TMP oleate polyol ester is more susceptible to hydrolysis than phthal-
ate diester. The three triaryl phosphate esters gave very low change in acid number,
easily meeting the ISO 10050 requirement for turbine control fluid service.
Copper corrosion rates were all acceptable and equivalent within the repeatabil-
ity range. The POE1 and PAG fluids gave the highest level of insolubles.
The three phosphate ester fluids all easily meet the unused fluid electrical resis-
tivity requirements of ISO 10050. The PAG fluid, although not a phosphate ester
but also marketed as an EHC fluid, fails to meet this requirement.

Anti-Wear/Lubricity Bench Test Data


Some common bench tests that can be used to compare anti-wear characteristics of
lubricants include the Falex Pin and Vee, the Four-Ball, and Cameron–Plint tests.
These tests were run for the eight non-aqueous hydraulic fluids chosen for this
study. The test results are discussed in the following sections.

FALEX PIN-AND-VEE BLOCK METHOD


The Falex Pin and Vee test (ASTM 3233-93 Method A [10]) measures the extreme
pressure properties of fluid lubricants.
Test description (wording based on the Summary of Test Method section of the
ASTM method):
A steel journal (pin) is rotated at 290 6 10 rpm against two stationary V-blocks
immersed in the lubricant sample. Load is applied to the V-blocks by a ratchet
mechanism. Method A increases the load continually (up to 4500 lb) until failure.
The maximum load is the point corresponding to the maximum torque applied
before either breakage or deformation of the specimen, causing a dramatic drop in
torque. At the failure point, either the pin breaks or deforms sufficiently to not
allow the ratchet mechanism to take up the next higher load.
Table 4 and Fig. 1 show the test results for the eight fluids. The test repeatability
is 627 % of the mean.
Fluids PHE2 and PHE3, both butylated phenyl phosphate esters, gave the best
results with maximum failure loads of 3010 lb. Interestingly, the PHE1 fluid (xylen-
ated phosphate ester) did not perform as well as the butylated phenyl phosphates.
REA AND BARKER, DOI:10.1520/STP157320130101 163

TABLE 4 Falex Pin-and-Vee data (ASTM D3233-93 Method A).

Coded Fluid Maximum Maximum Maximum Maximum Temperature at Failure


Name Torque, in-lb Load, lb Friction Coefficient Failure,  C Mode

PHE1 102.9 1152 0.27 78.8 Break


PHE2 116 3010 0.11 156.8 Deform
PHE3 109.8 3010 0.11 153 Deform
POE1 60.5 1332 0.13 85.3 Break
POE2 61.5 1106 0.17 82.8 Break
PAG 81.9 1479 0.16 88.7 Break
PAOE 138.3 714 0.58 83.7 Break
MO 12.9 176 0.22 83.4 Deform

The mineral oil fluid, MO, gave the poorest performance with a failure load of only
176 lb. Differences in additization may explain some of these differences.

FOUR-BALL METHOD
The Four-Ball test (ASTM D4172-94 [11]) measures the wear preventive character-
istics of lubricating fluids.
Test description (wording based on the Summary of Test Method section of the
ASTM method):
Three steel balls are clamped together and covered with lubricant sample. A
fourth steel ball (called the top ball) is pressed with a force into the cavity above the
three clamped balls, this providing three-point contact. The temperature of the
lubricant is controlled at a desired value, while the top ball is rotated at a desired
speed. Lubricants are compared for anti-wear characteristic by using the average
size of the scar diameters worn on the three clamped balls.

FIG. 1 Falex Pin-and-Vee maximum load (ASTM D3233-93 Method A).


164 STP 1573 On Fire Resistant Fluids

TABLE 5 Four-Ball wear data (ASTM D4172-94). Test conditions: 75 C, 1200 rpm, 40 kg, 1 hour

Coded Fluid Name Average Ball Wear Scar, mm

PHE1 0.56
PHE2 0.58
PHE3 0.56
POE1 0.53
POE2 0.66
PAG 0.44
PAOE 0.51
MO 0.85

Table 5 and Fig. 2 show the average wear scars for the eight fluids based on
duplicate runs (average of twelve scar measurements, six per run, two per stationary
ball). The test conditions are shown on Table 5 (75 C, 1200 rpm, 40 kg, 1 h). The
test repeatability is 60.12 mm.
The PAG fluid gave the best results, with an average wear scar of 0.44 mm. The
MO fluid gave the poorest results with an average wear scar of 0.85 mm. The POE2
fluid gave a result of 0.66 mm. The other fluids gave similar results in the approxi-
mate range of 0.50 to 0.55 mm.

CAMERON–PLINT WEAR TEST (BALL-ON-PLATE)


The Cameron–Plint TE77 machine is a tribometer test machine. It is a low fre-
quency, long displacement reciprocating sliding wear tester that can be run in dif-
ferent modes.

FIG. 2 Four-Ball average wear scar (ASTM D4172-94).


REA AND BARKER, DOI:10.1520/STP157320130101 165

FIG. 3 Cameron–Plint TE-77 schematic (ball-on-plate).

Figure 3 shows a schematic of the Cameron–Plint TE-77 tribo tester when used
in “ball-on-plate” mode. Test conditions are also shown in Fig. 3.
In this test mode, a steel spherical ball is rubbed at 30 Hz against a hardened
steel plate under a constant load of 100 N. The temperature is ramped up from
room temperature, holding at 50 C for 15 min, then at 100 C for 45 min, and
finally at 150 C for 15 min.
The average wear scars measured on the ball and plate are used to compare the
anti-wear characteristics of lubricants.
Table 6 and Fig. 4 shows the average ball wear scars for the eight fluids (single
runs, wear scars measured horizontally and vertically and averaged). The test
repeatability is 60.02 mm.

TABLE 6 Cameron–Plint TE-77 Ball-on-Plate test, ball wear data.

Coded Fluid Name Average Ball Wear Scar, mm

PHE1 0.51
PHE2 0.52
PHE3 0.52
POE1 0.56
POE2 0.62
PAG 0.34
PAOE 0.42
MO 0.64
166 STP 1573 On Fire Resistant Fluids

FIG. 4 Cameron–Plint average ball wear scar (ball-on-plate).

The Cameron–Plint ball scars correlate very well with the Four-Ball scar data
shown in Table 5 and Fig. 2. A comparison of Four-Ball and Cameron–Plint aver-
age ball wear scar data is shown in Fig. 5.
The PAG fluid gave the best Cameron–Plint results, with an average wear scar
of 0.34 mm. The next best fluid was PAOE with an average wear scar of 0.42 mm.
The MO and POE2 fluids gave the poorest results with average wear scars of
0.64 mm and 0.62 mm, respectively. The other fluids gave similar results in the ap-
proximate range of 0.50 to 0.55.
Table 7 and Fig. 6 show the maximum depth in micrometers of the plate wear
scars measured along the wear track on the plate.

FIG. 5 Comparative chart of four-ball and Cameron–Plint ball wear scar.


REA AND BARKER, DOI:10.1520/STP157320130101 167

TABLE 7 Cameron–Plint TE-77 Ball-on-Plate test, plate wear data.

Coded Fluid Name Maximum Plate Wear Scar Depth, micron

PHE1 29
PHE2 21
PHE3 23
POE1 25
POE2 34
PAG 31
PAOE 58
MO 207

The MO fluid gave a significantly deeper scar on the plate (207 lm) than all the
other fluids. The next worst fluid is PAOE (58 lm), while the remainder of the flu-
ids gave plate wear scar depths in the approximate 20 to 35 lm range.

CAMERON–PLINT FRICTION TEST (DOWEL-ON-PLATE)


The Cameron–Plint TE77 test machine can also be used to measure comparative
coefficients of friction when run in the “dowel-on-plate” mode. Figure 7 shows a
schematic of the Cameron–Plint TE-77 tribo tester when run in this mode. Test
conditions are also shown in Fig. 7.
In this test mode, a steel cylindrical dowel is rubbed at 5 Hz against a steel hard-
ened plate. The load and temperature are slowly increased to 100 N and 160 C,
respectively. Coefficient of friction data are dynamically captured at 100 N over the
temperature range.
Figure 8 shows the friction coefficients measured for the eight fluids over the
temperature range of 60 to 160 C.

FIG. 6 Cameron–Plint maximum plate wear scar depth (ball-on-plate).


168 STP 1573 On Fire Resistant Fluids

FIG. 7 Cameron–Plint TE-77 schematic (dowel-on-plate).

The highest VI fluids (POE1, POE2, PAG) give the lowest coefficients of fric-
tion over most of the temperature range, in the 0.06 to 0.085 range. POE2 gave the
lowest coefficient over the entire temperature range, in the 0.06 to 0.07 range. The
two fluids (PAOE and MO) with intermediate VI values give intermediate coeffi-
cients in the 0.085 to 0.10 range.
The lowest VI fluids (PHE1, PHE2, PH3) give the highest coefficients of friction
over most of the temperature range. This suggests a film effect is in play for these
low VI fluids, with significantly thinner lubricant films for the phosphate fluids,
especially at temperatures > 100 C. Note, however, that the two butylated phos-
phates PHE2 and PHE3 give similar and significantly higher coefficients (in the 0.09
to 0.14 range) than for the xylenated phosphate PHE1 (in the 0.075 to 0.12 range).

FIG. 8 Cameron–Plint coefficient of friction data (dowel-on-plate).


REA AND BARKER, DOI:10.1520/STP157320130101 169

SUMMARY OF ANTI-WEAR/LUBRICITY BENCH TESTS


The Falex Pin and Vee test results show the PHE2 and PHE3 butylated phenyl
phosphate esters to be significantly better than all the other fluids. This perform-
ance may be the result of the formation of a protective iron polyphosphate film.
However, the xylenated phosphate fluid (PHE1) did not perform as well. The min-
eral oil MO fluid gave the worst Falex performance.
The Four-Ball and Cameron–Plint “ball-on-plate” tests show the polyalkylene
glycol PAG fluid and polyalphaolefin diester PAOE fluid to be significantly better
than all the other fluids.
The mineral oil MO fluid gave the worst Four-Ball and Cameron–Plint
performance.
The Cameron–Plint “dowel-on-plate” test results show the highest VI fluids
(polyol ester POE1 and POE2 and polyalkylene glycol PAG) give the lowest coeffi-
cients of friction. The lowest VI fluids (phosphate esters) gave the highest coeffi-
cients, with the butylated phenyl phosphate PHE2 and PHE3 fluids having higher
coefficients than for the xylenated phosphate PHE1 fluid.
It should be restated that these performance measurements are for fully formu-
lated oils and so include the effects of both base stock and additives.
Overall, taking all these bench test data into account, the tested synthetics are
judged to be superior to the tested mineral oil for anti-wear/lubricity performance.
Amongst the tested synthetics, however, it is more difficult to clearly rank the dif-
ferent fluid types given the mixed results of the chosen bench tests. For example,
POE2 gave the lowest Cameron–Plint coefficient of friction, but relatively high
Four-Ball and Cameron–Plint ball wear results. As such, additive effects may be
influencing the ball wear results for POE2.
Depending on the bench test chosen, one synthetic type may appear to be bet-
ter than another. This may or may not translate to actual field performance. For
example, the suppliers of the tested phosphate esters, polyalkylene glycol, and pol-
yol esters all claim passing wear results in the Vickers V104C vane pump test.
Regardless of which bench test is used as a pre-screener, an end user should ulti-
mately rely on an appropriate hydraulic pump test to verify the anti-wear/lubricity
performance of a specific synthetic type.

Fire Resistance Bench Testing


FLASH AND FIRE POINTS
ASTM D92-12 [12] describes the determination of flash point and fire point by a
Cleveland open cup apparatus. This is a bulk ignition test. Test repeatability is
68 C.
Test description (wording based on the Summary of Test Method section of the
ASTM method):
An open flame is applied to the vapor space above a pool of liquid. The flash
point is the lowest fluid temperature at which application of the test flame causes
170 STP 1573 On Fire Resistant Fluids

TABLE 8 Flash point and fire point data (ASTM D92-12).

Coded Fluid Name Flash Point,  C Fire Point,  C

PHE1 262 356


PHE2 266 366
PHE3 264 362
POE1 274 370
POE2 322 382
PAG 273 308
PAOE 254 288
MO 222 240

the vapors above the test fluid to ignite. The fire point is determined by continuing
to heat the test fluid until application of the test flame causes the test fluid to ignite
and sustain combustion for a minimum of 5 s.
Table 8 and Fig. 9 show the flash points and fire points for the eight fluids.
The polyol ester fluids POE1 and POE2 have the highest fire points (370 and

382 C), followed closely by the triaryl phosphate esters PHE1, PHE2, and PHE3
(356 to 366 C). The lowest fire point is for the mineral MO fluid (240 C).
Based on fire point data alone, one might predict that the fire resistance of the
polyol ester fluids and the triaryl phosphate ester fluids are comparable. However,
as will be seen in the following sections, fire point data can be misleading for pur-
poses of ranking fire resistance.

FIG. 9 Flash point and fire point data (ASTM D92-12).


REA AND BARKER, DOI:10.1520/STP157320130101 171

FIG. 10 “Hot Manifold” test (ISO 20823) apparatus photograph.

FIRE RESISTANT BENCH TESTS


ISO 12922 requires hydraulic fluids to be tested against a set of fire resistance bench
tests that vary the ignition mode and the physical form of the fluid. These tests are
summarized here:
• ISO 20823 (hot surface; stream of fluid and pooled fluid)
• ISO 14935 (open flame; fluid adsorbed on substrate)
• ISO 15029-1/-2 (open flame; spray of fluid)

These tests can be used to predict the ignition and combustion behavior of flu-
ids in different fire potential scenarios. The following sections describe each test
and review the test results.

“HOT MANIFOLD TEST” (ISO 20823)


Figure 10 shows a photograph of the test apparatus.
Test description (wording taken from the Principle section of the ISO test
method):
In this test, a 10 ml portion of fluid is dropped from a predetermined height
and at a specified rate onto a tube heated to 700 C or another temperature in a se-
ries. The resulting spray from impact on the hot tube is examined for flash or burn,
both on the tube and after dripping from the tube and pooling in a catch tray below
the tube. Temperature varies across the length of the manifold tube and is hotter in
the center. Reported ignition temperatures are measured in the center of the tube.
Table 9 and Fig. 11 show the hot manifold ignition temperatures for the tested
hydraulic fluids. There is no precision data available for this test method. It is left to
each laboratory to produce a quality control procedure based on fluids of known
performance against which other fluids are ranked (Ref. [3]).
172 STP 1573 On Fire Resistant Fluids

TABLE 9 “Hot Manifold” test data (ISO 20823).

Coded Fluid Name Ignition Temperature,  C Flaming Drips to Tray? Flames in Tray?

PHE1 741 No No
PHE2 726 No No
PHE3 710 No No
POE1 509 Yes No
POE2 495 Yes Yes
PAG 458 Yes Yes
PAOE 474 Yes Yes
MO 444 Yes Yes

The triaryl phosphate esters give the best fire resistance performance as meas-
ured by this test. These fluids gave the highest ignition temperatures, ranging from
710 to 741 C.
The other four synthetic fluids and the mineral oil fluid all gave significantly
lower ignition temperatures (509 C). The mineral oil gave the lowest ignition
temperature (444 C).
Table 9 also shows whether or not flaming drips were observed falling to the
catch tray below, and whether or not any flames were observed in the catch tray
below.
The three triaryl phosphate esters flashed on the tube at the ignition tempera-
ture but did not produce any flaming drips. The drippings of all the other fluids
continued to burn on the way down to the catch tray. There was no burning in the
catch tray for any of the triaryl phosphate esters or for the POE1 fluid. The other
fluids (POE2, PAG, PAOE, and MO) continued to burn in the catch tray.

FIG. 11 “Hot Manifold” test (ISO 20823) ignition temperatures.


REA AND BARKER, DOI:10.1520/STP157320130101 173

FIG. 12 “Wick Test” (ISO 14935) apparatus photograph.

“WICK FLAME PERSISTENCE TEST” (ISO 14935)


Figure 12 shows a photograph of the test apparatus.
Test description (wording taken from the Principle section of the ISO test
method):
In this Pass/Fail test, a length of non-flammable aluminosilicate board is soaked
in the fluid being tested and placed in a reservoir of the fluid with an exposed edge.
A small flame is applied to the exposed edge of the board, and the persistence is
measured, in seconds, of the flame after removal of the igniting flame.
A total of six determinations is carried out for each of five different periods of
flame application (2, 5, 10, 20, and 30 s). The persistence of these five different peri-
ods of application of the igniting flame are calculated, and the result is the largest of
these averages. If the fluid continues to burn for >60 s, then the fluid fails the test.
Table 10 and Fig. 13 show the Pass/Fail results and the average burn times for
the 30 s flame application. There is no precision statement available for this test
method [4].
The triaryl phosphate esters PHE1, PHE2, and PHE3 gave the best fire resist-
ance performance, as measured by this test. These were the only fluids that passed
the test.

“SPRAY FLAME PERSISTENCE TEST” (ISO 15029-1)


Figure 14 shows a photograph of the test apparatus.
Test description (wording taken from the Principle section of the ISO test
method):
In this Pass/Fail test, a sample of fluid is pressurized and heated to a specific
pressure and temperature, and then atomized through a defined nozzle. The spray
174 STP 1573 On Fire Resistant Fluids

TABLE 10 “Wick Flame Persistence” test data (ISO 14935).

Coded Fluid Name Pass < 60 s Fail > 60 s 30 s Flame Application Burn Time, s

PHE1 Pass 2.2


PHE2 Pass 1.2
PHE3 Pass 2.7
POE1 Fail >60
POE2 Fail >60
PAG Fail >60
PAOE Fail >60
MO Fail >60

FIG. 13 “Wick Flame Persistence” test data (ISO 14935).

FIG. 14 “Spray Flame Persistence” test (ISO 15029-1) apparatus photograph.


REA AND BARKER, DOI:10.1520/STP157320130101 175

TABLE 11 “Spray Flame Persistence” test data (ISO 15029-1).

Coded Fluid Name Pass < 30 s Fail > 30 s Maximum Burn Time, s

PHE1 Pass 3
PHE2 Pass 8
PHE3 Pass 2
POE1 Fail 125
POE2 Fail 125
PAG Fail 134
PAOE Fail 124
MO Fail 130

produced is ignited with an oxyacetylene test flame of specific energy at various


points along the whole length of the spray pattern.
After ignition, the test flame is withdrawn, and the time is measured that the
flame continues to burn. The test result is the maximum unsupported burning
time. Flames burning <30 s are a Pass; flames burning >30 s are a Fail.
Table 11 and Fig. 15 show the Pass/Fail results and the maximum burn times.
There is no precision statement available for this test method [5].
The triaryl phosphate esters PHE1, PHE2, and PHE3 gave the best fire resist-
ance performance as measured by this test. These were the only fluids tested that
passed the test.

“STABILIZED SPRAY FLAME TEST” (ISO 15029-2)


Figure 16 shows a photograph of the test apparatus.
Test description (wording taken from the Principle section of the ISO test
method):
A pre-conditioned flux of test fluid is delivered to a test chamber through a
twin-fluid atomizer. Compressed air, supplied to the nozzle at a controlled rate, is

FIG. 15 “Spray Flame Persistence” test data (ISO 15029-1).


176 STP 1573 On Fire Resistant Fluids

FIG. 16 “Stabilized Spray Flame” test (ISO 15029-2) apparatus photographs.

used to produce an atomized spray, which is exposed to a defined flame of a gas


burner present throughout the test.
The gas flame acts to produce, by input of heat at a steady rate, a stabilized
spray flame, so that combustion properties, such as the rate of energy release and
flame length, are sufficiently steady over time to allow time-averaged values to be
measured.
Calculations of functions, such as ignitability factor (RI), flame length index
(RL), and smoke density (D) are made with these measurements.
Ignitability factor (RI) is a calculated value based on relative flame temperature
rise for the gas flame alone and for the gas flame with the test fluid (in both cases
comparing air inlet temperature and gas exhaust temperature). Flame length index
(RL) is a normalized flame length, numerically equal to 5000 mm divided by the
stabilized flame length in millimeters. Smoke density (D) compares light obscura-
tion before and after test fluid flow.
There is no precision statement for this test method [6]. However, each labora-
tory must run a series of ethylene glycol/water calibration fluids for ignitability
REA AND BARKER, DOI:10.1520/STP157320130101 177

TABLE 12 “Stabilized Spray Flame” test grading system (per ISO 15029-2). (Fire resistance
decreases going from category A to H)

Category A B C D E F G H

Ignitability Factor, RI >100 100–80 79–65 64–50 49–36 35–25 24–14  13


Flame Length Index, RL >100 100–56 55–51 50–11 10–7 6 — —
Smoke Density, D <0.01 0.01–0.05 0.051–0.1 < 0.1 — — — —

factor. The calibration fluids range from 10 to 100 % ethylene glycol. The raw test
fluid RI values are then corrected for the calibration results.
A classification scheme for the performance of fire resistant fluids is suggested
by ISO 15029-2. The primary reference is ignitability factor (RI), with secondary
factors being flame length index (RL) and smoke density (D). Table 12 summarizes
the classification system. The most fire resistant fluid would be graded as “A/A/A”
for RI/RL/D. The least fire resistant fluid would be graded as “H/F/D.” Aqueous flu-
ids rank high on this scale of fire resistance (towards “A/A/A”).
Table 13 shows the classification results for the eight non-aqueous hydraulic
test fluids. Note that only ignitability factor (RI) and flame length index (RL) were
measured for this study.
The triaryl phosphate esters overall give the best fire resistance performance as
measured by this test, giving ratings of E–F for ignitability factor and D for flame
length index. All the non-phosphate ester fluids received a lower flame length index
rating of E.
For ignitability factor, the PAG fluid did better than all the other non-
phosphate fluids with an RI rating of G (albeit a borderline H rating). The POE1,
POE2, PAOE, and MO fluids all received the lowest RI rating of H.

TABLE 13 “Stabilized Spray Flame” test results (ISO 15029-2).

Coded Fluid Ignitability Ignitability Fire Flame Length Flame Length Fire
Name Factor RI Resistant Category Index RL Resistant Category

PHE1 40 E 14 D
PHE2 29 F 11 D
PHE3 41 E 14 D
POE1 6 H 8 E
POE2 5 H 8 E
PAG 14 G 8 E
PAOE 4 H 7 E
MO 4 H 7 E
178 STP 1573 On Fire Resistant Fluids

SUMMARY OF FIRE RESISTANCE BENCH TESTS


The triaryl phosphate ester fluids overall gave consistently superior results to the other
fluids in the fire resistance tests required by ISO 12922 for HFDR and HFDU fluids.
The superior fire resistance of triaryl phosphate esters makes this type of non-
aqueous hydraulic fluid preferred in applications where ignition and combustion
risk for the fluid is high. For example, the fire risk is high in power plants running
modern steam turbines where superheated steam line temperatures can reach
600 C [13]. A leak of triaryl phosphate ester fluid onto a 600 C steam line will
not result in fluid ignition (see the “hot manifold” test results shown on Table 9 and
Fig. 11). However, the probability for fluid ignition would be high for any of the
other tested fluid types.
Another real world example is the situation of a fluid leak under piping insula-
tion in a power plant fire where the insulation is exposed to an open flame from the
fire. The “wick flame persistence” test results shown in Table 10 and Fig. 13 are ap-
plicable to this situation. Triaryl phosphate ester fluid soaking the insulation would
only burn for a very brief time before self-extinguishing. Any of the other tested
fluid types would continue to burn once ignited.
The triaryl phosphate ester fluids are difficult to ignite due in part to their stable
aromatic ring structure. In addition, once ignited, these fluids are self-extinguishing
due to low heats of combustion and self-quenching endothermic reactions in the
flame, which are characteristic of phosphorus combustion chemistry (Refs. [14,15],
p. 95 and p. 617, respectively).
The superior fire resistance performance of the triaryl phosphate esters com-
pared to other non-aqueous hydraulic fluids is consistent with earlier work as docu-
mented in Refs. [14–17].

Conclusions
This study compared the fire resistance and other key performance properties of
eight commercially available non-aqueous hydraulic fluids. These fluids are fully
formulated (additzed) fluids. Five base stocks were included in this study: triaryl
phosphate ester, polyol ester, polyalkylene glycol, polyalphaolefin and diester blend,
and mineral oil.
The requirements shown in ISO 12922 [2] and ISO 10050 [8] should be used as
the minimum guidelines when selecting a hydraulic fluid. The comparative data gave a
range of results for physical and chemical properties, air and water separability, and
hydrolytic stability. In some cases, the relevant ISO standard requirement was not met.
The synthetic hydraulic fluids tested are superior to the tested mineral oil for
anti-wear/lubricity performance. However, ranking the different synthetics for anti-
wear/lubricity performance is difficult based solely on the chosen bench tests:
• Butylated phenyl phosphates PHE2 and PHE3 gave the highest Falex fail loads
• PAG and PAOE fluids gave the lowest Four-Ball and Cameron–Plint wear
scars
REA AND BARKER, DOI:10.1520/STP157320130101 179

• POE1, POE2 and PAG fluids gave the lowest Cameron–Plint friction
coefficients
The Cameron–Plint friction data also suggest that xylenated phosphate offers a
lubricity advantage compared to butylated phenyl phosphate.
Despite being outperformed in some of the chosen anti-wear/lubricity tests, tri-
aryl phosphates are well known in the industry as being effective anti-wear/lubricity
additives (Refs. [14], pp. 96–97). Rather than solely rely on bench tests, it is sug-
gested that the end user should ultimately rely on an appropriate hydraulic pump
test matched to the application in order to verify the anti-wear/lubricity perform-
ance of a specific synthetic type.
An assessment of a fluid’s fire resistance requires tests covering different hazard
scenarios [18]. These hazards can include fluid exposure to hot surfaces, fluid pool-
ing, exposure to open flames, fluid adsorption on a substrate, and fluid spraying.
This study used the fire resistance bench tests specified by ISO 12922. From a safety
standpoint, where fire resistance is critical to the application, triaryl phosphate
esters are realistically the only choice for a true fire-resistant hydraulic fluid. These
are the only fluids that demonstrate consistently high fire resistance performance
across a range of tests. These fluids are difficult to ignite and when ignited exhibit a
self-extinguishing behavior resulting from the base stock’s chemical structure, low
heat of combustion and the characteristics of phosphorus combustion chemistry.

ACKNOWLEDGMENTS
The writers would like to thank Chemtura Corporation, Petroleum Additives: Faith Corbo
and Frank DeBlase (anti-wear and friction bench testing); Bob Rowland (chemical struc-
ture drawings); Chemtura Europe, Petroleum Additives: Penny Norridge (consultant and
coordination of fire-resistant bench testing); Health and Safety Laboratory (Buxton, UK):
Stuart Jagger (fire-resistant testing and photographs of bench test equipment).

References

[1] ISO 6743-4, 1999, “Lubricants, Industrial Oils and Related Products (Class L)—Classifica-
tion—Part 4: Family H (Hydraulic Systems),” 2nd ed., International Standards Organiza-
tion, Geneva, Switzerland.

[2] ISO 12922, 2012, “Lubricants, Industrial Oils and Related Products (Class L)—Family H (Hy-
draulic Systems)—Specifications for Hydraulic Fluids in Categories HFAE, HFAS, HFB, HCC,
HFDR and HFDU,” 2nd ed., International Standards Organization, Geneva, Switzerland.

[3] ISO 20823, 2003, “Petroleum and Related Products—Determination of the Flammability
Characteristics of Fluids in Contact With Hot Surfaces—Manifold Ignition Test,” 1st ed.,
International Standards Organization, Geneva, Switzerland.

[4] ISO 14935, 1998, “Petroleum and Related Products—Determination of Wick Flame Per-
sistence of Fire-Resistant Fluids,” 1st ed., International Standards Organization, Geneva,
Switzerland.
180 STP 1573 On Fire Resistant Fluids

[5] ISO 10529-1, 1999, “Petroleum and Related Products—Determination of Spray Ignition
Characteristics of Fire-Resistant Fluids—Part 1: Spray Flame Persistence—Hollow-Cone
Nozzle Method,” 1st ed., International Standards Organization, Geneva, Switzerland.

[6] ISO 15029-2: Petroleum and Related Products – Determination of Spray ignition Charac-
teristics of Fire-Resistant Fluids – Part 2: Spray test – Stabilized Flame Heat Release
Method, International Standards Organization, First edition, 2012.

[7] Phillips, W. D., “The Electrochemical Erosion of Servo-Valves by Phosphate Ester Fire-
Resistant Hydraulic Fluids,” Lubr. Eng., Vol. 44, No. 9, 1988, pp. 758–767.

[8] ISO 10050, 2005, “Lubricants, Industrial Oils and Related Products (Class L)—Family T
(Turbines)—Specifications of Triaryl Phosphate Ester Turbine Control Fluids (Category
ISO-L-TCD),” 1st ed., International Standards Organization, Geneva, Switzerland.

[9] ASTM D2619-09: Standard Test Method for Hydrolytic Stability of Hydraulic Fluids (Bev-
erage Bottle Method), Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA, 2009.

[10] ASTM D3233-93: Standard Test Methods for Measurement of Extreme Pressure Proper-
ties of Fluid Lubricants (Falex Pin and Vee Block Methods), Annual Book of ASTM Stand-
ards, ASTM International, West Conshohocken, PA, 2009.

[11] ASTM D4172-94: Standard Test Method for Wear Preventive Characteristics of Lubricating
Fluid (Four-Ball Method), Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA, 2010.

[12] ASTM D92-12: Standard Test Method for Flash and Fire Points by Cleveland Open Cup
Tester, Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2012.

[13] Quinkertz, R., Ulma, A. Gobrecht, E., Wechsung, M., and Siemens A. G., “USC Steam Tur-
bine Technology for Maximum Efficiency and Operational Flexibility,” Proceedings of the
POWER-GEN Asia 2008, Kuala Lumpur, Malaysia, Oct 21–23, 2008, p. 5.

[14] Phillips, W. D., Placek, D. C., and Marino, M. P, “Neutral Phosphate Esters,” Synthetics, Min-
eral Oils, Bio-Based Lubricants—Chemistry and Technology, 2nd ed., L. Rudnick, Ed., CRC
Press, Boca Raton, FL, 2013, Chap. 4.

[15] Phillips, W. D., “Fire Resistance and Fire-Resistant Hydraulic Fluids,” Synthetics, Mineral
Oils, Bio-Based Lubricants—Chemistry and Technology, 2nd ed., L. Rudnick, Ed., CRC
Press, Boca Raton, FL, 2013, Chap. 37.

[16] Phillips, W. D., “The Use of Triaryl Phosphates as Fire-Resistant Lubricants for Steam
Turbines,” Lubr. Eng., Vol. 42, No. 4, 1986, pp. 228–235.

[17] Phillips, W. D., “A Comparison of Fire-resistant Hydraulic Fluids for Hazardous Industrial
Environments. Part 1. Fire resistance and lubrication properties,” J. Synth. Lubr., Vol. 14,
No. 3, 1997, 211–235.

[18] Jagger, S., Nicol, A., and Thyer, A., 2007, “Assessing Hydraulic Fluid Fire Resistance,” Ma-
chinery Lubrication, http://www.machinerylubrication.com/Read/1094/hyrdraulic-fluid-
fire-resistance (Last accessed 15 Mar 2013).
FIRE RESISTANT FLUIDS 181

STP 1573, 2014 / available online at www.astm.org / doi: 10.1520/STP157320140090

W. D. Phillips1

Assessing and Classifying the


Fire-Resistance of Industrial
Hydraulic Fluids: The Way
Ahead?
Reference
Phillips, W. D., “Assessing and Classifying the Fire-Resistance of Industrial Hydraulic Fluids:
The Way Ahead?,” Fire Resistant Fluids, STP 1573, John Sherman, Ed., pp. 181–198,
doi:10.1520/STP157320140090, ASTM International, West Conshohocken, PA 2014.2

ABSTRACT
The combustion behavior of hydraulic fluids can vary significantly but many
types are currently labeled as “fire-resistant.” This lack of discrimination cannot
only be confusing to the user, but may result in incorrect fluid selection with
potentially dangerous consequences. The paper therefore suggests an
alternative way of classifying ‘fire resistant’ fluids. However, for acceptance, the
methods used to measure “fire resistance” have to be agreed. The paper
therefore examines the principal methods in current use for determining this
property and proposes a test regime to help define the levels of performance.

Keywords
fire resistance, fire-resistant hydraulic fluids, ignitability, flame propagation,
self-extinguishment, phosphate esters, polyol esters, polyalkyleneglycols, spray
ignition, hot surface ignition, wick test

Introduction
Fire-resistant hydraulic fluids have two major functions to perform: to efficiently
transfer power and to provide “fire resistance.” While the technical requirements
for effective power transfer are well understood, the level of fire resistance

Manuscript received October 7, 2008; accepted for publication April 28, 2009; published online August 12,
2014.
1
Consultant, W David Phillips and Associates, 7 Kettleshulme Way, Poynton, Stockport, Cheshire SK12 1TB,
United Kingdom.
2
ASTM Symposium on Fire Resistant Fluids on June 24, 2013 in Montreal, Quebec, Canada.

Copyright V
C 2014 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
182 STP 1573 On Fire Resistant Fluids

(or safety) that is offered by these fluids is often less well appreciated. This is at a
time when equipment operating conditions continue to increase in severity in the
search for higher operating efficiency and reduced equipment costs while, concur-
rently, there is greater emphasis on ensuring high levels of health and safety at work.
Additional factors such as fluid costs, and concerns regarding the effect on
health and the environment of some existing fluid types, have resulted in different
fluid chemistries being introduced as “fire-resistant” hydraulic fluids, sometimes on
the basis of very limited evidence. Unfortunately, the end user does not necessarily
understand that fire resistance can vary significantly between fluid types or the limi-
tations of some of the tests often quoted as measuring fire resistance. In selecting a
fire-resistant fluid it is essential that the user can readily identify the level of safety
offered by the fluid and can match this to the hazards of the application.
The purpose of this paper is therefore to examine how fire resistance is cur-
rently assessed; the variation in fluid performance that is found under these test
conditions and how, in the future, we can better describe the combustion behavior
in terms of the level of fluid performance.

What is a Fire-Resistant Fluid?


The term fire resistance is perhaps most widely used in the building industry where
it is defined as

“The ability of a material, product, assembly or structure to fulfill for a stated


period of time the required stability, integrity, thermal insulation, and/or other
expected duty specified in a standard fire-resistance test” (ISO 13943 [1]).
In the tests specified for the building industry, the resistance of the product to a
standard (and usually fairly constant) source of heat is measured. When applied to
a fluid, the terminology has a somewhat different meaning. The emphasis being less
on time of resistance and more on the temperature required for ignition. While
building materials are unlikely to display any significant change in properties during
testing, fluids will evaporate (which can result in a change in composition), and will
probably degrade thermally and oxidatively. When the temperature is high enough,
oxidation in the vapor phase will result in combustion. With the exception of water
(or fluids containing a high percentage of water), all current fluid types will combust
in the presence of a high energy source applied for a sufficient period of time.
The available definitions of a fire-resistant fluid (not necessarily a hydraulic
fluid) are not very helpful. For example, ASTM D4175-09, “Terminology Relating
to Petroleum, Petroleum Products and Lubricants,” [2], defines a fire-resistant fluid
as, “Any liquid that is able to withstand fire or give protection from fire,” while ISO
5598: Fluid Power Systems–Vocabulary [3] is more specific, “Fluid difficult to ignite
which shows little tendency to propagate flame.”
Unfortunately these definitions are too vague for use with hydraulic fluids. The
ASTM definition fails to define “fire:” to indicate for how long the fluid must
PHILLIPS, DOI 10.1520/STP157320140090 183

“withstand fire” and under what conditions. It also fails to address the consequences
of ignition or combustion, which are almost as important as ignitability itself.
The ISO definition, while an improvement, is still unsatisfactory as it does not
clarify how difficult the fluid must be to ignite. There is also no guidance on what is
meant by “little tendency to propagate flame” or the conditions under which this is
to be measured.
The use of the term ‘fire-resistant’ for a large group of fluids can therefore mask
a potentially wide variation in their performance. The loose wording of these defini-
tions is one reason why so many fluids are being marketed under this description
and why, in today’s litigious and safety-conscious climate, it is important to be
more precise.
In the hydraulic fluid industry, fire-resistance is also regarded as relative to that
of mineral oil. However, the fire resistance of mineral oil will vary depending on its
viscosity and the refining process involved in its manufacture. Both parameters will
affect the content of low molecular-weight hydrocarbons which are the most flam-
mable components. It therefore seems sensible to specify the viscosity grade and
API Group of oil when it is used as a reference for providing comparative data with
fire-resistant fluids. Preferably the mineral oil should be of the same viscosity as the
test fluid. Some fire test data on mineral and hydrocarbon oils of different types and
viscosity are given in Table 1.
A new definition of a fire-resistant hydraulic fluid is therefore urgently required
and will be considered later. It should relate to both ignitability and flame propaga-
tion tendency under standard test conditions with mineral oil used as a reference.

Fire-Resistant Hydraulic Fluid Types


These fluids are currently divided into two different types depending on how their
fire-resistance is achieved.
(1) Fluids where water provides the fire resistance. These types may contain any-
thing from 20 % w/w water (so-called HFCE-fluids) to 99 % w/w water
(HFA fluid types). The classification for industrial hydraulic fluids (HFA/
HFC etc.) is according to ISO 6743–4 [4].

TABLE 1 Fire resistance data on refined mineral oils and a synthetic hydrocarbon in comparison
with some fire-resistant hydraulic fluids.

Autoignition Manifold Ignition


Flash Point ( C) Fire Point ( C) Temperature ( C) Temperature ( C)
Product ISO VG ASTM D92-05 ASTM D92-05 ASTM D2155-66 ISO 20823

Min. Oil-API Gp. II 32 234 262 400 370


Min. Oil-API Gp. III 46 251 282 385 395
PAO-API Gp. IV 46 258 286 410 390
HFDU polyol ester 68 256 340 425 430
HFDR phosphate 46 260 350 585 750
184 STP 1573 On Fire Resistant Fluids

(2) Fluids which possess fire-resistance as a result of their chemical composition.


These are mainly non-aqueous products falling into the HFDR (phosphate
esters) and HFDU classification (polyol esters, polyalkyleneglycols, and vege-
table oils).
Water-based fluids are difficult to test and to compare with non-aqueous fluids.
Their fire-resistance depends principally on the water content. However, while the
fire resistance will be very good under hazard conditions where the contact time
with the ignition source is short (e.g., spray tests), where prolonged contact with the
ignition source occurs (e.g., hot surface ignition), it is possible for the water to evap-
orate and ignition to take place. Fluids with a low water content (e.g., HFCE types)
are obviously at a greater risk of this occurring.
The delay in ignition with these fluids has been suggested as “escape time” [5]
and promoted as a potential advantage for these fluids. However, it is important to
appreciate that the fluid can still ignite and the delay may or may not be advanta-
geous depending on the circumstances of the release (e.g., when the rate of temper-
ature increase is high the delay may be very short and not significantly different to
that for a non-aqueous fluid).
With non-aqueous hydraulic fluids the situation is different. For these fluid
types, undoubtedly the most important property is their ease of ignition. However,
the behavior of the fluid if and when it combusts is also important. For example, if
much heat is released by ignition, it will pose a hazard to personnel in the vicinity
and influence the rate of growth of the fire. Conversely those fluids which have a low
heat release (or which self-extinguish when not in contact with the ignition source)
could produce a lower rate of fire spread [6]. Therefore, tests evaluating the behavior
of the burning fluid, particularly its tendency to propagate flame, are also necessary.
HFDU fluids rely on their lower volatility/higher molecular weight to raise their
ignition temperature but may also contain a polymeric thickener to increase the
droplet size in a spray. This makes the spray more difficult to ignite and may assist
in improving the performance of the fresh fluid under this specific test condition
but, as the polymer shears down in use, the fire-resistance trends to that of the
polymer-free product. It also fails to improve performance under other test condi-
tions, e.g., hot surface ignition.
The fire resistance of HFDR fluids (aryl phosphate esters) arises from a combi-
nation of physical and chemical properties (thermal/oxidative stability) and the fact
that some of the vapor-phase reactions are endothermic (heat absorbing) resulting
in reduced flame propagation. As a consequence non-aqueous “HFD” fluids have a
wide range of fire-resistance which is confusing to the user and could cause a fluid
with inadequate fire resistance to be used with resulting safety concerns.

Classification and Selection of Test Methods


The most important aspects of the combustion behavior of a fluid are as
follows:
PHILLIPS, DOI 10.1520/STP157320140090 185

(1) the ease of ignition or ignitability of the fluid;


(2) the behavior of the fluid on ignition (i.e., does it propagate flame or is it self-
extinguishing).
As a consequence, the test methods that are available for assessing combustion
behavior fall into such categories as follows:
• ease of ignition/ignitability;
• heat release or flame propagation;
• smoke and other products of combustion, or both; and
• an intrinsic physical property (e.g., heat of combustion).

However, many tests claim to measure more than one parameter and therefore
this classification is unsatisfactory. An alternative scheme might be to categorize
them according to the physical form of the fluid or the mode of ignition:
• bulk fluid ignition;
• ignition of a pool or thin film;
• ignition of a droplet, intermittent stream, spray, or jet;
• ignition when absorbed on a substrate;
• ignition by an open flame;
• ignition by a hot surface; or
• ignition by a spark.

In practice, combinations of the above approaches have been used but the main
emphasis has been on selecting tests which simulate the potential application haz-
ards. Where this is done the test conditions should simulate or reflect that hazard
as closely as possible. As stated in BS 6336 [7]:

“If more than one hazard is present then different tests may be required for
each hazard. In practice, simplifications may be necessary to reduce the time
and cost involved but the conditions finally adopted should relate as nearly as
possible to the actual environment in which the hazard is thought to arise. If a
test is designed to assess different products for a specific use it is essential that
the method be capable of being applied to all possible materials, and on an
equal basis (author’s emphasis).”
When selecting tests to represent a specific hazard, preference should be given
to the following:
• Standard tests, i.e., those issued by a national or international standards
organization.
• Tests which have precision data. In the absence of precision data a reference
or calibration fluid(s) which are made from pure components may be used.
• Tests which discriminate between fluids, i.e., are able to rank fluids in order of

fire resistance. A pass/fail test is of limited use as there is no indication of how


good (or bad) the fluids are relative to one another (or to any reference fluid).
However, even when all these restrictions are applied it should be acknowledged
that, as with any laboratory test, it is only possible to assess fluid behavior under very
specific conditions. Although these are supposed to be representative of service, such
is the variety of operating conditions and possible fluid release mechanisms in use
186 STP 1573 On Fire Resistant Fluids

that an accurate prediction of the mode of fluid escape (and hence an accurate simu-
lation) is usually impossible to accomplish. But perhaps this is not essential. In most
cases only a comparison of performance under the same conditions, either between
different fluids or with a reference fluid, is sought. It should also be remembered that
fluids can behave differently in different tests and that, as stated in BS 6336 [7], in
order to obtain a complete picture of fire resistance behavior, performance in several
different tests is usually required. This is essential when, as is frequently the case, sev-
eral different hazards may exist for the same application.
One assumption made in fire-resistance testing is that no change in perform-
ance under test (or in use) can take place. Unfortunately there are conditions when
this can occur as follows:
– If water is lost when testing water-based products.
– If polymeric materials shear down in use resulting in a reduction in viscosity.
– If the material degrades thermally or oxidatively in use to form significant
amounts of flammable degradation products.
– Contamination of the operating fluid with mineral oil or another more com-
bustible material.
Test selection is therefore a complex procedure. This is not helped by the large
number of tests which purport to measure some aspect of fluid fire-resistance. In a
relatively recent count there were over 50 [8]! This probably reflects the difficulty of
translating the wide variation in conditions under which combustion can take place
into a simple and precise test method.
Table 2 identifies the most widely used tests for assessing hydraulic fluid fire re-
sistance and which can be used to assess a range of different fluid types. Not all are
capable of ranking fluids in their current form.

TABLE 2 Standard fire tests for industrial hydraulic fluids.

Physical Form of Fluid

Bulk Ignition (Pool Droplet, Stream, Adsorbed on


Ignition Mode and Thin Film) Spray, or Jet Substrate

Flash and Fire Wick Tests:


Open flame Points: “Spray” tests: ASTM D5306-92
(piloted ASTM D92 ISO 15029-1/2 ISO 14935
ignition) ISO 2592 Factory Mutual 6930 Std. IEC 1197
Manifold ignition:
ISO 20823
Autoignition:
Hot surface ASTM D2155/E 659-78
Spark NA NA NA

Note: NA ¼ none available.


PHILLIPS, DOI 10.1520/STP157320140090 187

There are, of course, other types of standard fire test (e.g., oxygen index, cone calo-
rimetry, etc.) which, for various reasons, are not specified for hydraulic fluids. Further
information on these procedures and their limitations can be found in Refs 8 and 9.
It will be seen in the table that there are no current tests for assessing fire-
resistance in the presence of a spark. This may reflect the low probability of such a
form of ignition and also the difficulty in selecting test conditions which could be
regarded as “representative.”
Current practice in evaluating and comparing fire resistance therefore focuses
on the procedures given in Table 3. Of these the spray tests, autoignition, manifold
ignition, and the ISO wick test are the most widely used. Although not a hazard
related test, use is sometimes made of net heat of combustion measurements in
assessing fluid fire-resistance, for example, in the Factory Mutual Corp. spray igni-
tion test. More detailed comments on this test are given later under the assessment
of the Factory Mutual procedure.

The Performance of Different Fluid Types in


Standard Fire Resistance Tests
The data given in Table 3 are based on fluids that are commercially available and are
a mixture of information available in product data sheets and the results of tests
that have been carried out by independent laboratories. Obviously it is impossible
to provide data on all the different fluids that are commercially available but it is
believed that the figures quoted are representative of their respective types. Further
information on fluid performance is available in Refs 6, 8, and 10.
No data are presented on fluid types HFAE/HFAS as these fluids normally con-
tain such a high water content that they are very difficult or impossible to ignite
and, in any case, most of the standard tests cannot be used to evaluate these fluids.
As can be seen from the table:
• Flash and fire point values do not correlate with other test data.
• With the exception of HFDR fluids, most fluids display similar autoignition
temperature values including the water-glycol fluid.
• A much wider spread of results is seen with the manifold ignition test [11].
• Autoignition [12] and manifold ignition test data for mineral oil and HFDU
fluids are fairly similar.
• The ISO 15029–2 [13] spray test discriminates well between the fluids,
whereas the Factory Mutual test [14] shows less discrimination and in some
cases its results conflict with the ISO rating. This is mainly because of the
incorporation of a “critical heat flux” measurement (see below).
• All the HFDU fluids propagated flame in the Wick test [15].

Use of Current Test Methodology to Assess


Fluid Fire Resistance
There are two main specifications for industrial fire-resistant fluids, each with a dif-
ferent approach to assessing fire resistance. The first is ISO Standard 12922 [16]
188
STP 1573 On Fire Resistant Fluids
TABLE 3 Typical fire-resistance performance for the different fluid types.

Fluid Typea

Fire-Resistance Test Min Oil HFBb HFCb HFCE HFDRb HFDUb (polyol esters) HFDUb (veg. oils) HFDUb (PAG)

Flash point ( C) 230–250 NRc NRc NRc 240–270 260–320 >290 270
- ASTM D92
Fire point ( C) 260–285 NRc NRc NRc 335–370 335–378 >345 310
– ASTM D92
Autoignition ( C) 380–400 385 425 No 545– > 625 400–435 390 395
– ASTM D2155 data
Manifold ignition ( C) 370–395 >650 >650 >650 700–750 440–480 420–465 400–420
– ISO 20823
Spray ignition-ignitability class H F B-C E D-E G-H H F
– ISO/DIS 15029-2
Spray ignition-spray flamm. factor Not classified Not tested Approved No data Approved Approved/ Approved/ Approved/
specification specification specification
testedd Testedd testedd
– FM 6930 std.
Wick flame persistence (secs) Continuous Pass Pass No data Pass Continuous burning
burning
– ISO 14935
a
The data are typical for fluids of ISO VG 46 and 68 viscosity grades;
b
HFA, HFB, etc., are the ISO definitions of different types of hydraulic fluids and taken from ISO standard 6743–4 [4].
c
NR¼Not relevant.
d
“Specification tested” means that the fluid failed to meet the FM Approved Fluid limits.
PHILLIPS, DOI 10.1520/STP157320140090 189

which covers all grades of fluid used in general industrial applications and the sec-
ond is Factory Mutual Standard “Flammability Classification of Industrial Fluids,
Class Number 6930” [14].
ISO 12922, Lubricants, industrial oils and related products (class L)—Family H
(hydraulic systems)—Specifications for categories HFAE, HFAS, HFB, HFC,
HFDR, and HFDU; this standard uses three separate ISO fire-resistance tests:
– ISO 15029: Spray test (This standard is currently in three parts but two of the
methods–Parts 1 and 3–are likely to be of limited interest in the future and
most attention will focus on Part 2, the Stabilized Flame Heat Release
Method [13];
– ISO 20823: Manifold ignition test; and
– ISO 14935: Wick flame persistence method.
Each of these methods is capable of evaluating a wide range of fire-
resistant fluids though it is unnecessary to evaluate the fire-resistance of prod-
ucts with a very high water content (>80 %). ISO 15029–2 and ISO 20823 are
capable of ranking the different fluids while ISO 14935 is currently a pass/fail
procedure. The first two procedures primarily assess ignitability while 14935
(and also 20823) evaluate flame propagation. The advantage of using three tests
means that we have confirmation of the performance under ignitability and
propagating conditions and an attempt to modify the performance of the fluid
by the use of additives (e.g., adding droplet modifiers to reduce ignitability in a
spray) may be identified as they do not significantly improve hot surface igni-
tion performance.
ISO/DIS 15029-2 was developed in order to harmonize EU test requirements
for fire-resistant fluids and to be able to evaluate all the different types of fluid [10].
It is an “ease of flame stabilization” test and measures exhaust gas temperatures
of the burning fluid. This test is based on the concept that the more easily a fluid
combusts, the greater the amount of heat is released.
The exhaust temperature of the burning test fluid is compared with the temper-
ature without the test fluid but with the igniter operating. From these measure-
ments an ignitability factor (IF) is obtained. However, to rank fluids (and to take
into consideration the potential repeatability of the test) the range of IF values is
broken down into “Classes,” from A to H, with A being the least flammable and H
the most flammable. Figure 1 shows a schematic of the test equipment while the
results of testing some commercially available fluids are shown in Fig. 2 [17]. Full
details of the method are, of course, to be found in the ISO Standard while the back-
ground to its development can be found in Reference 18.
Advantages:
• The test can be adapted to measure heat release rates or oxygen depletion but,
for simplicity, temperature measurements are preferred. It can also measure
other combustion parameters such as smoke production.
• Although no formal precision data are yet available (see below) a series of
calibration (or reference) fluids based on mixtures of distilled water and pure
ethylene glycol are used to ensure the consistency of the test conditions. Some
190 STP 1573 On Fire Resistant Fluids

FIG. 1 The ISO/DIS 15029-2 Spray Flammability Test apparatus.

attempts are made to control the ambient environment in terms of tempera-


ture and humidity as these have been shown to be highly influential in deter-
mining the test result [19], however further restrictions may be required.
• This test is able to evaluate and compare most of the different types of fire-
resistant fluid. However, fluids containing >80 % water will not normally
combust under these test conditions.

FIG. 2 Results of the ISO/DIS 15029-2 Spray Test on a range of fire-resistant hydraulic
fluids.
PHILLIPS, DOI 10.1520/STP157320140090 191

Disadvantages and concerns:


• The test “does not give a total measure of the heat output as it does not mea-
sure radiative heat but uses convective heat as a means of providing a relative
indication of the degree of completeness of combustion” [18].
• The current procedure uses volume flow rates and therefore does not take into
consideration the significant difference in density between fluids. Without
such compensation the comparison between products is not strictly valid.
• This is a relatively expensive test to carry out requiring relatively large volumes
of fluid, and presently there are only four test rigs known to be in existence.
Currently no precision test data (repeatability or reproducibility) are available
but following the publication of the standard test method (expected in 2009) a
limited round robin test program is to be initiated.
ISO 20823 is a method that previously appeared in Aeronautical Material
Standard 3150 C and measures the ignitability of fluid when in contact with a hot
surface; in this case the hot surface is in the form of a manifold or tube.
Advantages:
• This is a relatively low cost test that is able to compare most fluid types which
ignite up to a temperature of about 800 C. It requires only small amounts of
test fluid.
• The test also examines the propensity to propagate flame after the fluid has
moved away from the ignition source.
Disadvantages and concerns:
• Currently there is no precision data but the use of reference fluids is being
examined together with the possibility of comparative testing on the existing
test rigs.
• Since residence time is a key variable with hot surface tests, the ignition tem-
peratures would not be expected to be identical with tests carried out on a flat
surface but the ranking of fluids would be expected to be unchanged.
ISO 14935 examines the flammability of a fluid when soaked onto an adsorbent
“wick.” A propane flame is applied to the top edge of a piece of ceramic board for
varying times and the time it takes for any resulting flame to self-extinguish is
measured.
Advantages:
• This is an easy, low-cost test which requires only small amounts of fluid.
• It is capable of testing and comparing most fluids except those with a very
high water content.
Disadvantages and concerns:
• The test is of the pass/fail type and therefore does not lend itself to the produc-
tion of precision data.
• It does not currently measure the relative tendency of the flame to propagate
although it could be adapted to do so.
The Factory Mutual Standard 6930 (January 2002)-Flammability Classification
of Industrial Fluids is a method that was developed to replace a previous spray test
that measured the persistence of burning but was a pass/fail procedure and suscep-
tible to spray droplet size. A new method was therefore required to rank all the
192 STP 1573 On Fire Resistant Fluids

different fluid types and, if possible, to eliminate the effect of polymeric thickeners
on spray flammability [20]. It was eventually decided to use a heat release test utiliz-
ing some of the components from the previous spray ignition test but the new test
involved a vertical burn so that the products of combustion could be more easily
measured [20]. Figure 3 [14] shows a schematic of the test equipment.
Instead of relying on measurements of exhaust gas temperatures as in the ISO
15029-2 method, the FM procedure measures the generation rates of CO and CO2
and calculates the chemical heat release rate from these. In the early stages of the
test development, it was found that volatile (and flammable) solvents such as meth-
anol, ethanol, and heptane gave low heat release values. In order to avoid such

FIG. 3 The Factory Mutual heat release measurement apparatus.


PHILLIPS, DOI 10.1520/STP157320140090 193

products being confused with fire-resistant hydraulic fluids, the concept of “critical
heat flux for ignition” was introduced [21]. An expression for a dimensionless spray
flammability parameter (SFP) was derived by combining the total chemical heat
release data with the critical heat flux in the following equation:

(1) SFPnormalized ¼ 11:02  106 Qch =qf qcr mf

where:
Qch ¼ chemical heat release rate (kW),
qf ¼ density of the fluid (kg/m3), and
qcr ¼ critical heat flux for ignition (kW/m2)
and

(2) qcr ¼ a  r  T 4

where:
a ¼ fluid surface resistivity (assume to be unity),
r ¼ Stefan-Boltzmann constant (5.67  1011 kW/m2  K4),
T ¼ fire point temperature ( K), and
mf ¼ fluid mass flow rate during the heat release rate measurement (g/s).
Advantages:
• Both the spray and fire point tests use relatively small amounts of fluid.
• The spray test can be used to assess smoke production and also flame length.

Disadvantages and concerns:


• The test is not a national or international standard (but is known
internationally).
• The test is expensive and requires sophisticated equipment.
• As a result of problems in the testing of water-based fluids, the above proce-
dure is no longer applicable to fluids of type HFA/B/C and a different test pro-
tocol for these fluids is currently under development. Therefore there can be
no comparison between water-containing and non-aqueous fluids. Fluids
with 60 % water (w/w) are regarded as possessing excellent fire resistance
and testing of these products is therefore not required [21].
• The ability to rank fluids is limited to two broad categories by SFP limits; these
are “FM approved” and “FM approvals specification tested.” Approved fluids
require no additional fire protection over that specified for construction or
occupancy of a building while “specification tested” products may require
additional fire protection due to the hazard posed by the fluid.
• Currently there is only one test facility and therefore full precision data
(repeatability and reproducibility) cannot be established. No repeatability data
appear to have been published.
• The test only investigates the ignitability behavior of the fluid. It does not take
into consideration the behavior of the combusting fluid once it is removed
from the ignition source, i.e., its propagating tendency.
194 STP 1573 On Fire Resistant Fluids

• No attempt has been made to control ambient test conditions (e.g., air temper-
ature and humidity), which have been shown elsewhere to have a significant
effect on test results.
• The calculation of chemical heat release rate assumes that the heat released by
the production of CO/CO2 is solely responsible for the heat released and that
no other reactions are involved. While this assumption might be valid for
products containing C, H, and O, it is not safe to assume this applies to prod-
ucts containing, for example, phosphorus. On combustion, phosphorus com-
pounds will also produce phosphorus oxides which will react quickly with any
water present to produce phosphorus acids, therefore the combustion reaction
mechanism is significantly more complicated than the sole production of
oxides of carbon. All these other reactions will influence the net heat released.
• “The use of CO /CO2 generation as a way of calculating heat release rates is
regarded as substantially less accurate than measurement of O2 consumption
as ‘fuels’ do not have a universal constant for this correlation” [22].
• As was mentioned above, the concept of critical heat flux to ignition was intro-
duced in order to compensate for the volatility of some fluids. After examining
several alternative tests it was eventually decided to use fire point (open cup)
and combine both heat release rate and critical heat flux into one mathemati-
cal expression called the Spray Flammability Parameter. The justification for
combining the two terms is not clear as volatile liquids could have easily been
identified and eliminated by limits on separate tests but, as a result, the SFP
value is significantly influenced by the fire point result as its value (in the
equation above) is present to the fourth power. The method states thata1 %
error in the measurement of fire point results in a critical heat flux error of
>2 %. Since the repeatability value for open cup fire point is 8 C, for a fluid
with a fire point of 350–400 C (typical for non-aqueous fluids) the error in
critical heat flux determination could therefore be >4 %. This would be in
addition to any errors in the measurement of the mass of CO/CO2. The results
of fire point tests (a piloted ignition test) do not necessarily correlate with
those of other fire tests. For example, trioctyl phosphate has a fire point of
170 C but a manifold ignition temperature of 465 C (cf. polyol ester val-
ues of about 350 C and 450 C, respectively).
• The FM procedure suggests that net heat of combustion measurements by
ASTM D240, can be used in the absence of chemical heat release rate data.
• This is of concern for the following reasons:
• ASTM D240-02, “Standard Test Method for Heat of Combustion of Liquid
Hydrocarbon Fuels by Bomb Calorimetry”, is currently only valid for products
containing C, H, O, S, and N [23]. It is not valid for products containing other
elements, e.g., phosphorus and therefore cannot be used to compare the com-
plete spectrum of fire-resistant fluids.
• ASTM D240 assumes the products of combustion are carbon dioxide, nitrogen
oxides, sulfur dioxide, and water. If, say, a phosphate ester is combusted, phos-
phorus oxides would be expected to be produced and these oxides react rap-
idly with any water present to produce phosphorus acids. (Water is present in
the test fluid as an impurity; is generated as an oxidation product from the
PHILLIPS, DOI 10.1520/STP157320140090 195

TABLE 4 Fire test behavior on different fire-resistant fluid types (ISO VG 46 Fluids) (see Ref. 24).

Fire Resistance Test Results

Spray Ignition Manifold Factory Mutual


Test-Ignitability Ignition Wick Flame 6930 Std. –Spray
Factor Test  C Persistence Flammability
Fluid (ISO/DIS15029-2) (ISO 20823) (ISO 14935) Test Status

Mineral oil Class H 380 Fail Not tested (probable fail)


Polyalkyleneglycol Class F 420 Fail Approved
Vegetable ester Class H 420 Fail Approved
Polyol ester (1) Class H 440 Fail Approved
Polyol ester (2) Class G 455 Not tested Not tested
Polyol ester (3) Class H 435 Fail Specification tested
Phosphate ester Class E >704 Pass Approved
Water-glycol fluid Class A >704 Pass Approved

hydrocarbyl portion of the molecule and, if the procedure is followed as speci-


fied, a small amount is actually added to the test fluid in the bomb). Thus any
heat released may involve not only the heat of combustion but also the heat of
formation (or heat of reaction) with water and therefore the data are not com-
parable with heat of combustion from other fluids falling within the scope of
this method.
• Some chemicals, e.g., phosphate esters are very difficult to combust completely
by conventional bomb calorimetry. Successful methods use very small quanti-
ties (5 ll) of fluid, e.g., by microcalorimetry. There is no known evidence to
confirm that phosphate esters show “complete” combustion by the ASTM
D240 procedure which involves substantially greater amounts (0.5–1.0 g).
It should be apparent from the above that neither approach to the measure-
ment of the fire resistance of hydraulic fluids is completely satisfactory and in fact
they produce conflicting results. Table 4 shows the effect of testing the same prod-
ucts by the different protocols [24].

Can the Conflict be Resolved?


Obviously the conflict between the two standards needs to be resolved and agree-
ment reached on the tests used to assess fluid fire-resistance. In the absence of such
an agreement, end users may select a product that does not have the level of fire-
resistance necessary for their application.

A Change in Terminology?
As indicated earlier, one aspect that must be seriously considered is whether the
continual use of the term fire-resistant can be justified for all fluid types. This, in
itself, is misleading. In the past the use of terms “fire-retardant” and “less flamma-
ble” have been used for some fluid types, for example by Factory Mutual Research,
196 STP 1573 On Fire Resistant Fluids

and the latest revision of ISO 12922 refers to the HFDU Class of fluids as less flam-
mable. In reality, the only fluids that can justify the term fire-resistant are the high
water-based fluids where the water content is >80 %. One possibility would
therefore be to divide the fluids into the following classes:
– Fire resistant: (Fluids which, under standard test conditions, are extremely dif-
ficult or impossible to ignite and do not propagate flame.) This definition
would cover mainly ISO classes HFAE and HFAS (where the water content
is >80 %)
– Fire retardant: (Fluids which, under standard test conditions, are difficult to
ignite and do not significantly propagate flame.) This definition would cover
ISO classes HFB, HFC, HFDR, and HFAE/S fluids where the water content
is <80 %
– Less flammable: (Fluids which, under standard test conditions, are more diffi-
cult to ignite than mineral oil but readily propagate flame.) This would cover
HFDU fluids.
While this would represent an improvement over the current situation, we are
still left with some uncertainty as to what is meant by “more difficult to ignite,” etc.
A further difficulty could be the translation of these terms into other languages
while retaining the same nuances. Therefore some quantification of this terminol-
ogy becomes essential.
It might also be thought that the continued use of the term fire-resistant could
cause confusion and that its replacement by, for example, “non-combustible” might
be preferred.

Defining the Terminology by Performance


To assume that a certain fluid chemistry must always have the same combustion
behavior would be to deny both the possibility of future developments and the cur-
rent reality. A better approach would be to define the performance of each level of
performance by standard test methods. Unfortunately the FM test is not sufficiently
discriminating but it is possible with the ISO standards. A proposed definition of
each of the three categories above using performance in the ISO 12922 tests is given
in Table 5. In order to be classified as fire resistant or less flammable, etc., it would

TABLE 5 A proposed classification of fluid fire test behavior by description and performance.

Spray Ignition Manifold


Test-Ignitability Ignition Wick Flame
Fluid Factor—Minimum Test  C Persistence sec.
Description (ISO 15029-2 Class) (ISO 20823) (ISO 14935)

Fire-resistant B No ignition No Ignition


Fire-retardant E >550 <60
Less flammable G 400–550 Report
Mineral oil (for comparison) H 380 Fail
PHILLIPS, DOI 10.1520/STP157320140090 197

be necessary to meet all three limits. This would avoid the possibility of using addi-
tives which were effective under one test condition only.

Conclusion
The use of the description fire-resistant for an increasingly wide range of industrial
hydraulic fluids is unsatisfactory and potentially dangerous as it might suggest that
all fluids have a similar level of performance, whereas a considerable variation
exists. Changing the terminology (e.g., by classifying different levels of perform-
ance) will help but the different categories will have to be defined by performance
in suitable tests. Unfortunately, no single test can describe the behavior under all
potential hazard conditions and therefore several tests will be required to assess
both ignitability and flame propagation behavior.
The group of fire tests currently listed in ISO 12922, although not yet meeting
all the criteria for suitable procedures, offers the potential for such a test regime. It
is suggested that this combination of testing (with appropriate limits) and descrip-
tions similar to those given above for the different levels of performance are consid-
ered for future use.

References
[1] ISO 13943, “Fire Safety–Vocabulary,” 2000.

[2] ASTM D4175-05: “Standard Terminology Relating to Petroleum, Petroleum Products,


and Lubricants,” Annual Book of ASTM Standards, ASTM International, West Consho-
hocken, PA, 2005.

[3] ISO 5598, “Fluid Power Systems and Components–Vocabulary,” 2008.

[4] ISO 6743-4, “Lubricants, Industrial Oils and Related Products (Class L)-Classification–
Part 4: Family H (Hydraulic Systems),” 1999.

[5] McDonald, W. F., “Escape Time–An Important New Concept in Fire Resistance,” The
Houghton Line, Vol. 34, 1970, E. F. Houghton Co., Valley Forge, PA, pp. 1–12 and pp. 41–43.

[6] Jagger, S. F., Nicol, A. N., and Thyer, A. M. “A Comprehensive Approach to the Assess-
ment of Fire-Resistant Hydraulic Fluid Safety,” UK Health & Safety Laboratory Report
No. FR/2002/05, 2005.

[7] British Standard 6336, “Development and Presentation of Fire Tests and Their Use in
Hazard Assessment,” 1998.

[8] Phillips, W. D., “Fire-Resistance tests for Fluids and Lubricants–Their Limitations and
Misapplication,” Fire Resistance of Industrial Fluids, ASTM STP 1284, 1996, ASTM Interna-
tional, West Conshohocken, PA.

[9] Jagger, S. F. and Thyer, A. M., “A Review of Test Methods to Assess the Fire Hazards of
Fire-Resistant Hydraulic Fluids,” UK Health and Safety Laboratory Report No. FR/04/
05, 2005.
198 STP 1573 On Fire Resistant Fluids

[10] Phillips, W. D., “Spray Ignition Testing of Hydraulic Fluids-Options and Uncertainties,”
Proceedings of the National Conference on Fluid Power, Las Vegas, March, 2005.

[11] ISO 20823, “Petroleum and Related Products–Determination of the Flammability Char-
acteristics of Fluids in Contact with Hot Surfaces–Manifold Ignition Test,” 2003.

[12] ASTM D2155-63: “Autoignition Temperature of Liquid Petroleum Products,” Annual


Book of ASTM Standards, ASTM International, West Conshohocken, PA, 1963.

[13] ISO/DIS 15029-2, “Petroleum and Related Products–Determination of Spray Ignition


Characteristics of Fire-Resistant Fluids–Part 2: Spray Test–Stabilized Flame Heat Release
Method,” 2008.

[14] Factory Mutual Research Corp., “Flammability Classification of Industrial Fluids, Class
6930,” Factory Mutual Corp. Approval Standard, 2002.

[15] ISO 14935, “Petroleum and Related Products–Determination of Wick Flame Persistence
of Fire-Resistant Fluids,” 1998.

[16] ISO 12922, “Lubricants, Industrial Oils and Related Products (Class L) –Family H (Hy-
draulic Systems) –Specifications for Categories HFAE, HFAS, HFB, HFC, HFDR and
HFDU,” 2008.

[17] Data provided by the Health and Safety Executive, Buxton, UK and the Laboratorie Cen-
trale de Houlliéres du Bassin de Lorraine, Marienau, France.

[18] Yule, A. J., and Moodie, K., “A Method for Testing the Flammability of Sprays of Hydrau-
lic Fluid,” Fire Saf. J., Vol. 18, 1992, pp. 273–302.

[19] Jagger, S. F., Macmillan, A. J. R., and Allen, J. T., “Reproducibility and Repeatability of
Results of Hydraulic Fluids from the New Buxton Spray Test,” UK Govt. Health and
Safety Executive, Research and Laboratory Services Division Report No. IR/L/FR/91/15,
1991.

[20] Khan, M. M., “Spray Flammability of Hydraulic Fluids and Development of a Test Meth-
od,” Factory Mutual Research Technical Report No. J.I. OTOW3.RC, May 1991.

[21] Daday, G., “FM Approvals Standard for Industrial Fluids,” Proceedings of the STLE An-
nual Meeting, Philadelphia, May 6–10, 2007.

[22] Babrauskas, V., personal communication, June 17, 2008.

[23] ASTM D240-02: “Heat of Combustion of Liquid Hydrocarbon Fuels by Bomb Calo-
rimeter,” Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA,
2002.

[24] Anglin, J., “Fire-Resistant Fluids for Aluminium Processing,” Proceedings of the STLE An-
nual Conference, Cleveland, May 17–22, 2008.

You might also like