Petroleum Science and Technology
Petroleum Science and Technology
Samuel Hsu · Paul R. Robinson
Petroleum
Science and
Technology
Petroleum Science and Technology
Chang Samuel Hsu Paul R. Robinson
•
Petroleum Science
and Technology
123
Chang Samuel Hsu Paul R. Robinson
Petro Bio Oil Consulting Katy Institute for Sustainable Energy
Tallahassee, USA Katy, TX, USA
Department of Chemical and Biomedical
Engineering
Florida A&M University/Florida State
University
Tallahassee, FL, USA
China University of Petroleum—Beijing
Changping, Beijing, China
This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Driven largely by economic growth in the developing world, the use of petroleum
continues to expand. Use of renewable fuels and chemicals is expanding even more
rapidly, but petroleum will remain important for decades to come. We wrote this
book with the hope that it will help to educate students and professionals who are
striving to exploit this valuable resource efficiently, cleanly, and safely.
Our first book together is Practical Advances in Petroleum Processing, which
appeared in 2013. Practical Advances was followed 4 years later by a more com-
prehensive work: Springer Handbook of Petroleum Technology. The Handbook
expanded the content to include upstream exploration (discovery) and production
(recovery). Its success is demonstrated by the fact that it was downloaded over
80,000 times within 16 months after publication, according to Bookmeatrix.1
Chapters were contributed by leading authorities from integrated oil companies,
catalyst suppliers, licensors, consulting ventures, and academic researchers around
the globe.
The present book has a different focus. It is designed to serve as a textbook and
reference. It includes areas not discussed in the previous books: midstream oper-
ations and petroleum-derived chemicals.
The contents are based on materials used for the course Petroleum Science and
Technology, which has been taught at Florida A&M University/Florida State
University since 2012. It is divided into 18 chapters to cover fundamentals,
upstream exploration and production, downstream refining for fuels and lubricants,
petrochemicals and their derivatives, and mid-stream operations. In the last
chapter, regulations for safe operations and environmental protection are described
along with examples of incidents from throughout the energy and chemical
industries.
1
http://www.bookmetrix.com/detail_full/book/9c19f1e2-b01d-4a26-b277-2a487a16570b#citations.
v
vi Preface
We wish to thank all of the chapter authors of our previous two books. Most
of all, we wish to thank our devoted, magnificent wives, Grace Miao-Miao Chen
陳妙妙 and Carrie Robinson, for putting up with our absences—mental if not
physical—during so many nights and lost weekends throughout the preparation of
this book.
vii
viii Contents
xvii
xviii List of Figures
Fig. 7.6 Arrangement of a steering tool, orienting sub, and bent sub
for directional drilling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 104
Fig. 7.7 Main offshore oil production regions in the world
as of 2012 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 105
Fig. 7.8 Mobile offshore drilling units . . . . . . . . . . . . . . . . . . . . . . . .. 105
Fig. 7.9 Major offshore production structures and systems . . . . . . . . .. 107
Fig. 7.10 A colossal offshore platform lights up the night off
the coast of Norway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 108
Fig. 7.11 Primary oil recoveries by gas drive and water drive . . . . . . .. 108
Fig. 7.12 Generic depiction of injection and production wells
for secondary and tertiary recovery . . . . . . . . . . . . . . . . . . . .. 109
Fig. 7.13 Recovery by alkaline-surfactant-polymer (ASP) flooding. . . .. 112
Fig. 7.14 Huff and puff cyclic steam injection . . . . . . . . . . . . . . . . . . .. 113
Fig. 7.15 Steam drive injection (steam flooding or steam
stimulation) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Fig. 7.16 Toe-to-heel air injection (THAI) for oil recovery. . . . . . . . . . . 115
Fig. 7.17 Oil sand fields in northern Alberta, Canada . . . . . . . . . . . . . . . 118
Fig. 7.18 Steam assisted gravity drainage (SAGD) . . . . . . . . . . . . . . . . . 120
Fig. 7.19 Hydraulic fracturing (fracking) . . . . . . . . . . . . . . . . . . . . . . . . 121
Fig. 8.1 Distillation equipment described by 3rd Century Greek
alchemist Zosimos of Panopolis. . . . . . . . . . . . . . . . . . . . . . .. 130
Fig. 8.2 Selected world events and crude oil prices, 1859–2015 . . . . .. 131
Fig. 8.3 The 3:2:1 crack spread for Brent crude between June 2012
and May 2013 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 136
Fig. 8.4 Industry-average volume-percent yields from petroleum
refineries in the United States . . . . . . . . . . . . . . . . . . . . . . . . . 137
Fig. 8.5 A topping refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Fig. 8.6 A hydroskimming refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
Fig. 8.7 A catalytic cracking refinery . . . . . . . . . . . . . . . . . . . . . . . . . . 148
Fig. 8.8 A coking refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Fig. 8.9 Integrated refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Fig. 8.10 Integration of refinery and petrochemical plant . . . . . . . . . . . . 151
Fig. 8.11 Oil refinery process units: upgrading processes
for top-of-the-barrel cuts—VGO and lighter—which
boiling below about 1050 °F (556 °C) . . . . . . . . . . . . . . . . .. 152
Fig. 8.12 Oil refinery conversion processes for bottom-of-the-barrel
cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 153
Fig. 8.13 Refinery product blending . . . . . . . . . . . . . . . . . . . . . . . . . . .. 156
Fig. 9.1 Flow diagram of desalting involving an electrostatic
desalter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Fig. 9.2 Cross-section view of a desalter . . . . . . . . . . . . . . . . . . . . . . . 162
Fig. 9.3 Single- and Two-Stage electrostatic desalting systems . . . . . . . 163
Fig. 9.4 Simplified diagram of crude distillation . . . . . . . . . . . . . . . . . . 164
Fig. 9.5 Overlaps and tail ends of distillation cuts . . . . . . . . . . . . . . . . 165
xx List of Figures
xxv
xxvi List of Tables
Table 8.6 Ten U.S. states with the largest refining capacity . . . . . . . .. 140
Table 8.7 Ten largest refineries in the world (2018) . . . . . . . . . . . . . .. 140
Table 8.8 Refining capacity by country (in MBPCD:
thousand barrels per calendar day) . . . . . . . . . . . . . . . . . . .. 141
Table 8.9 Feeds and products for hydroprocessing units . . . . . . . . . . .. 143
Table 8.10 Molecular weight, H/C and boiling points for selected
hydrocarbons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Table 8.11 Equivalent distillation capacity index of refining units . . . . . 149
Table 9.1 Boiling ranges of distillation fractions . . . . . . . . . . . . . . . . . . 164
Table 9.2 Typical straight-run yields from various crudes . . . . . . . . . . . 171
Table 9.3 Average U.S. consumption of petroleum products,
1991–2003 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Table 9.4 Destinations for straight-run distillates . . . . . . . . . . . . . . . . . 173
Table 9.5 Feeds and products for hydroprocessing units . . . . . . . . . . . . 178
Table 10.1 Typical feed and product distributions in reforming . . . . . . . 194
Table 11.1 Zeolitic catalytic cracking versus cracking catalyzed
by an amorphous silica-alumina . . . . . . . . . . . . . . . . . . . . .. 217
Table 11.2 Representative FCC heat balance . . . . . . . . . . . . . . . . . . . .. 223
Table 11.3 Comparison between thermal cracking
and catalytic cracking . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 230
Table 11.4 Thermal versus catalytic cracking yields on similar
topped crude feed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 231
Table 11.5 Average bond lengths and energies of selected bonds . . . . .. 232
Table 11.6 Comparison of heats of reaction (kJ/mol) calculated
from bond energies with that heats of formation
for Reactions 11.1 and 11.2 . . . . . . . . . . . . . . . . . . . . . . . .. 233
Table 11.7 Hydrocracking product flexibility . . . . . . . . . . . . . . . . . . . .. 239
Table 12.1 Comparison of product yields for catalytic
and thermal cracking processes . . . . . . . . . . . . . . . . . . . . . . . 254
Table 13.1 Five API categories of lube base stocks . . . . . . . . . . . . . . . . 256
Table 13.2 Viscosity index of hydrocarbon types . . . . . . . . . . . . . . . . . . 258
Table 13.3 Typical solvents for extraction . . . . . . . . . . . . . . . . . . . . . . . 264
Table 13.4 Propane versus ketones dewaxing . . . . . . . . . . . . . . . . . . . . . 270
Table 13.5 Comparison of waxes from different sources. . . . . . . . . . . . . 276
Table 13.6 Properties of PAO derived from different linear
1-olefins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
Table 13.7 Basic properties of ester base stocks . . . . . . . . . . . . . . . . . . . 281
Table 13.8 Lubricating properties of selected PAGs . . . . . . . . . . . . . . . . 283
Table 15.1 Petroleum products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
Table 15.2 U.S. consumption of natural gas and petroleum products
(in 1000 barrels per day) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Table 15.3 ASTM specification numbers for hydrocarbon fuels . . . . . . . 306
Table 15.4 Properties of some gasoline blending components . . . . . . . . 313
Table 15.5 Additives used in gasoline . . . . . . . . . . . . . . . . . . . . . . . . . . 314
List of Tables xxvii
Natural gas in reservoirs contains mostly methane (CH4 ). Like crude oil, the origin
of natural gas is biological, and oil and gas are formed by the same natural forces.
Hence, most (but not all) petroleum reservoirs contain both oil and gas. We commonly
talk about “oil and gas” together. As ancient biomass is transformed into fossil
hydrocarbons, different oil- and gas-generation windows correspond to different
residence time at different depths. Methane is formed when liquid petroleum is
“over-matured” due to excess thermal stress in deep formations.
Methane can also be formed from the anaerobic decomposition of natural wet-
lands, rice paddies, emissions from livestock, organic wastes in landfills and biomass
burning (forest fires, charcoal combustion, etc.), as biogenic methane. The radioac-
tive carbon isotope, 14 C, is present in biogenic methane, but absent in fossil methane
in natural gas.
Natural gas that contains only traces of other compounds is dry gas. If natural gas
contains significant amounts of ethane, propane, butanes, and higher hydrocarbons,
it is called wet gas. The heavier components can be recovered individually or as con-
densate or natural gas liquid (NGL) in natural gas processing plants. Natural gasoline
is a condensate fraction comprised mostly of pentane. Sour gas contains hydrogen
sulfide, and acid gas contains carbon dioxide and/or hydrogen sulfide. Sour-gas pro-
cessing plants coproduce elemental sulfur, which is used to make sulfuric acid and
fertilizers. Some natural gas contains commercial quantities of inert gases—helium
(the product of α-decay in radioactive minerals underground), neon and/or argon.
Almost all commercial helium comes from natural gas plants as a byproduct [2].
1.2 Fossil Hydrocarbons (Fossil Fuels) 5
Methane is also present in another form as methane hydrate (clathrates) [3], where
methane is incarcerated in a cluster of water molecules as an ice-like material. Vast
amounts of methane hydrate deposits occur on the deep ocean floor on continental
margins and in places north of the arctic circle. It is estimated that methane hydrate
deposits contain around 6.4 trillion (6.4 × 1012 ) tonnes of methane [4]—twice as
much carbon as all other fossil fuels on earth. However, the necessary technol-
ogy for industrial production of the hydrates is not yet available. Hydrocarbon
clathrates cause problems for the petroleum industry, because they can form inside
gas pipelines, often resulting in obstructions.
The production of methane hydrate is fundamentally different than the extrac-
tion of oil and natural gas. The conventional recovery is based on the hydrocarbons
flowing naturally through the pores of reservoirs to the production well. Hydrates,
on the other hand, are solid. They must be dissociated first before methane can be
extracted. Several methods have been tested, such as hot water injection to break
down the hydrate and release methane, depressurization by drilling into the deposits
to release pressure for releasing methane, and carbon dioxide injection to replace
methane in the clathrate (molecular cage). Carbon dioxide hydrates are more stable
than methane hydrates. Carbon dioxide captured from natural gas and coal power
plants and injected into hydrates for storage is a strategy of carbon dioxde seques-
tration to reduce its emission into atmosphere. Each technique has its challenges
and limitations. Hence, which of these methods will be best suited for production at
industrial scales is still uncertain.
Crude oil is the common name for liquid petroleum. Crude oils, also called crudes,
are complex mixtures. There are hundreds of different crudes with significantly
different compositions. Crudes typically are named for their source country, reservoir,
and/or some distinguishing physical or chemical property. Table 1.1 presents selected
physical and chemical properties for ten crude oils.
The lightest liquid is condensate, which is essentially natural gas liquids, with
boiling points in the gasoline range. Light crude oils have low boiling points, low den-
sities (specific gravities), low viscosities, low sulfur, and low or negligible amounts
of nitrogen and other hetero-atom compounds. In sweet crudes such as Tapis (a
Malaysian crude), the sulfur content is low (0.028%). Sour crudes have more sulfur,
which gives them a tart taste. Synthetic crude oil is produced from coal, kerogen,
or natural bitumen. Processing costs are higher for conventional or synthetic crudes
with high density and large amounts of sulfur, nitrogen, and trace contaminants.
Shengli and many other Chinese crudes are very high in nitrogen, which can present
special challenges during processing.
6 1 Characteristics and Historical Events
Heavy crude oils possess high density—close to that of water. Extra Heavy
oils have densities greater than water. Both oils have very high boiling points and
high viscosities (>1000 centipoise (cP)). They also contain high concentrations of
heteroatom-atom containing hydrocarbons.
Distillation yields are an exceptionally important property of petroleum, because
distillation is the key step in separating crude oil into useful fractions, which deter-
mine the value of crude oil. Crudes containing larger amounts of light, low-boiling
fractions—naphtha, kerosene, and gas oil (diesel)—are more valuable. Table 1.2
shows distillation data for four common crudes. The naphtha content of Brent is
twice as high as Ratawi, and its vacuum residue content is 60% lower. Bonny Light
crude yields the most middle distillate and the least amount of vacuum residue.
Colloquially, the terms “bitumen,” “asphalt,” and “tar” are used interchangeably to
describe certain black, semi-solid mixtures of hydrocarbons. “Pitch” is an archaic
term for the same kind of substance. Geologists say “bitumen” when referring to
natural deposits, such as the famous La Brea Tar Pits. In the United States, bitumen
produced by crude oil refining is called “asphalt.” Outside the United States, bitu-
men from refineries is called “refined bitumen” or simply “bitumen”. The severely
degraded oil on oil sand, or tar sand, is also called “bitumen”.
Like extra heavy oils, natural bitumens have specific gravities greater than 1.0 (API
gravity <10), but the viscosities are higher (>10,000 cP). Under ambient conditions,
1.2 Fossil Hydrocarbons (Fossil Fuels) 7
Table 1.2 Distillation yields for four selected crude oils [5]
Source field Brent Bonny light Green canyon Ratawi
Country Norway Nigeria USA Mid East
API gravity 38.3 35.4 30.1 24.6
Specific gravity 0.8333 0.8478 0.8752 0.9065
Sulfur, wt% 0.37 0.14 2.00 3.90
Yields, wt% feed
Light ends 2.3 1.5 1.5 1.1
Light naphtha 6.3 3.9 2.8 2.8
Medium naphtha 14.4 14.4 8.5 8.0
Heavy naphtha 9.4 9.4 5.6 5.0
Kerosene 9.9 12.5 8.5 7.4
Atmospheric gas oil 15.1 21.6 14.1 10.6
Light VGO 17.6 20.7 18.3 17.2
Heavy VGO 12.7 10.5 14.6 15.0
Vacuum residue 12.3 5.5 26.1 32.9
Total naphtha 30.1 27.7 16.9 15.8
Total middle distillate 25.0 34.1 22.6 18.0
Naphtha plus distillate 55.1 61.8 39.5 33.8
natural bitumen is a soft and/or sticky solid, but when heated it flows. In practical
terms, it is recovered as a solid but transported and processed as a liquid by adding
diluents to lower viscosity. It is important to distinguish between natural bitumen and
refined bitumens. The latter are specialty products with rather tight specifications.
Refined bitumen is used primarily for paving and construction. Tar sands (also known
as oil sands) contain much of the world’s recoverable oil.
The largest bitumen/extra heavy oil deposits are in Venezuela with a total of 298
billion barrels. In 2nd place is Alberta, Canada, where the proven reserves are 173
billion barrels; in addition, the province holds 1.4 billion barrels of conventional
crude [6]. In comparison, for that same year, proven reserves of conventional crude
oil in Middle East were 804 billion barrels [7]. In tar sands, bitumen is associated with
sand and clay, from which it can be recovered with hot water or steam. Venezuela’s
oil sands are technically “extra heavy oil” deposits since they don’t contain bitumen.
The viscosities of Canadian tar sands vary widely, ranging from 10,000 to 600,000 cP,
while those of Venezuelan tar sands are more uniform, typically ranging from 4000
to 5000 cP [7]. In the United States, tar sands are found primarily in Eastern Utah,
mostly on public lands. These deposits contain 12 to 19 billion barrels of recoverable
oil.
8 1 Characteristics and Historical Events
1.2.4 Solids
Compared to liquids and gases, solids are harder to recover, transport and refine.
Liquids and gases can be pumped through pipelines and into refineries with rela-
tive ease. Slurries of coal and water can be transported as fluids, but the water must
be removed and eventually purified at considerable expense before the coal can be
burned or gasified. Solid coal is consumed on a large scale to produce heat, steam and
electricity. These days, coal-powered transportation vehicles are rare. Coal-burning
steam ships and railway locomotives are less efficient than their oil-powered coun-
terparts. Typically, the specific energy of petroleum is 90% greater than a ton of
bituminous coal and 40% greater than a ton of anthracite [8]. Even if for some
reason a railroad or shipping company wanted to burn coal, doing so wouldn’t be
practical due to the present lack of coaling stations. Coal is the most widely used
fuel in China and many other countries for power generation. Sulfur, ash, and trace
metals in the coal cause severe contamination of air with smog, acid rain, mercury,
and particulates. Upon combustion, the sulfur becomes sulfur oxides (SOx), primar-
ily SO2 . High-temperature combustion generates nitrogen oxides (NOx) from the
nitrogen and oxygen in air. Smog (photochemical smog) is generated by sunlight-
induced reactions between nitrogen oxides (NOx) and volatile organic hydrocarbons
(VOC); reaction products include ground-level ozone, an especially noxious pollu-
tant. Sulfate particulates, ranging in size from 1 to 20 microns, can be carried by
winds hundreds of miles, eventually returning to the earth as dry or wet “acid depo-
sition.” Wet deposits are commonly called “acid rain,” which also can contain NOx.
The combination of smog, particulates, and acid rain can be deadly, especially in
large cities.
Kerogen is the solid organic matter in sedimentary rocks. Unlike bitumen, it
doesn’t flow even when heated. But at high-enough temperatures—e.g., 900 °F
(480 °C)—it decomposes into gases, liquids, bitumen, and refractory coke. Huge
amounts of kerogen are trapped in oil shale deposits. Fenton et al. [9] estimated that
1.3 trillion barrels of shale oil could be recovered from the world’s oil shale reserves.
Table 1.3 presents composition information on Green River oil shale from the west-
ern United States. About 91% of the kerogen is hydrogen and carbon, but only 15%
of the shale is kerogen. Shale oil—synthetic crude from oil shale—tends to contain
high amounts of arsenic, a severe poison for refinery catalysts, and mercury. Usu-
ally, the arsenic is removed in existing hydrotreating units with special high-nickel
chemisorption catalysts, which trap the arsenic by forming nickel arsenides.
Coal is another non-petroleum hydrocarbon resource. It is a black or brown
combustible rock composed mostly of carbon, hydrocarbons and ash. Generally,
it is classified into four ranks—anthracite, bituminous, sub-bituminous, and lignite.
Anthracite is relatively rare, containing 86–97% carbon and has a high heating value.
Bituminous coal is far more common. It contains 45–86% carbon and is burned
to generate electricity. It is also used extensively in the steel and iron industries.
Sub-bituminous coal contains 35–45% carbon, and lignite contains 25–35% carbon.
Lignite is crumbly, has high moisture content and relatively low heating value. Over
1.2 Fossil Hydrocarbons (Fossil Fuels) 9
the years, special circumstances have driven the large-scale conversion of coal into
liquids, both directly and indirectly. Direct processes convert coal into various com-
binations of coal tar, oil, water vapor, gases, and char. The coal tar and oil can be
refined into high-quality liquid fuels [10].
Developed in 1925, the Fischer-Tropsch (F-T) process is the main indirect route
for converting coal into liquids. The coal is first gasified to make synthesis gas (syn
gas)—a balanced mixture of CO and hydrogen. Over F-T catalysts, synthesis gas is
converted into a full range of hydrocarbon products, including paraffins, alcohols,
naphtha, gas oils, and synthetic crude oil. The F-T process was used extensively in
Germany between 1934 and 1945. In South Africa, an improved version of the F-T
process is used on a large scale to manufacture chemicals and fuels.
Synthesis gas is also derived from natural gas via steam-methane reforming. It can
be converted into hydrogen and petrochemicals such as methanol. Worldwide, vast
amounts of hydrogen are used to produce ammonia via the Haber-Bosch process.
Examination of artifacts shows that humans were using petroleum long before writing
emerged as a means of conveying knowledge and recording events from one genera-
tion to the next. According to archaeologists, bitumen was used for hafting spears as
early as 70,000 BC near Umm el Tlel, in present-day Syria [11]. Neanderthals used
bitumen, too. A paper by Cârciumaru et al. provides evidence that Neanderthals in
10 1 Characteristics and Historical Events
Romania also hafted spears with bitumen between 28,000 and 33,000 BC; the dates
are based on uncalibrated 14 C dating [12, 13].
Jane McIntosh’s excellent book [14] about the ancient Indus valley shows that
baskets were water-proofed with bitumen before 5500 BC in Mehrgarh, an ancient
site located in present-day Pakistan between the cities of Quetta, Kalat and Sibi.
Bitumen is mentioned in some of the earliest records, specifically those written
on tablets in about 3200 BC and discovered in the ancient city of Sumer. Bitumen
use is also mentioned in Egyptian pictographs that were written at roughly the same
time. Sumer was the leading city of the Sumerian civilization, which arose in about
3500 BC in Mesopotamia—“the land between two rivers”—and lasted until about
1900 BC. The two rivers are the Tigris and Euphrates, located in present-day Iraq.
Sumerian writings describe the use of bitumen for mortar, to cement eyes into
carvings, for building roads, for caulking ships, and in other waterproofing applica-
tions. The asphalt came from nearby oil pits, and great quantities of it were found
on the banks of the river Issus, one of the tributaries of the Euphrates.
The Greek historian Herodotus mentioned the use of bitumen in Babylon (1900
to 1600 BC), including for construction of the famous Tower [15, 16].
From about the same time (3200 BC), Egyptian writings describe the use of pitch
to grease chariot wheels and asphalt in mummification, primarily to water-proof the
strips of cloth in which the mummies were wrapped. The Egyptians’ primary source
of bitumen was the Dead Sea, which the Romans called Palus Asphaltites (Asphalt
Lake).
As far back as 1500 BC, while drilling for brine, Chinese miners discovered natural
gas, which was used as a communal source for lighting and heating. The Chinese
were also drilling pioneers. Confucius wrote in 600 BC about using bamboo poles
to build pipelines and drill 100-foot natural gas wells.
In 347 AD, oil was being produced from bamboo-drilled wells in China. Bitumen
was slowly boiled to get rid of lighter fractions, leaving behind a thermoplastic
material with which scabbards and other items were covered. Statuettes of household
deities were cast with this type of material in Japan, and probably also in China.
Ancient Persian tablets tell about using bitumen and its fractions for lighting, top-
ical ointments, and flaming projectiles. It is likely that the light fractions were recov-
ered with simple batch distillation apparatus similar to those described by Zosimus,
an alchemist who lived at the end of the 3rd and beginning of the 4th century AD
[17]. By 500 BC, it was known that light fractions, such as naphtha could be used
not just for illumination, but also as a supplement to asphalt, making the latter easier
to handle.
Greek fire was invented during the reign of Constantine IV Pogonatus (668–685)
by Callinicus of Heliopolis, a Jewish refugee from Syria. This formidable incendiary
weapon was hurled onto enemy ships from siphons and burst into flame on contact
with air. It could not be extinguished with water. In 673 AD, Greek ships used the
weapon to defend Constantinople (today’s Istanbul, Turkey), crippling the Arab fleet
that was attacking the city [18]. The composition of Greek fire was kept as a top secret
and remains a matter of speculation and debate. However, Greek fire is believed to
be a viscous liquid composed of naphtha, liquid petroleum, bitumen and quicklime.
1.3 Use of Petroleum: A History 11
In more recent times, the French extracted oil from oil sands in the 1700s. In
the United States and Canada, oil appeared in brine wells and was recovered by
skimming.
The modern petroleum era began in the 1840s. In 1847, James Oakes built a “rock
oil” refinery in Jacksdale, England, to recover “paraffin oil” for lamps. Benjamin
Silliman Jr., a chemist hired by the Pennsylvania Oil Rock Company, determined
that 50% of a petroleum sample could be distilled into burning oils and 40% could
be employed for lubrication and gas lighting. The first modern oil well was drilled in
1848 by F. N. Semyenov in Azerbaijan. Eleven years later, an actual oil refinery was
constructed near the well to convert the raw materials into desired products. A large
milestone in petroleum products was reached when Canadian geologist Abraham
Gesner distilled kerosene from crude oil in 1848. The operation replaced the need
for whale oil for lamps and heating. The first true oil well in North America was
drilled in Petrolia, Ontario in 1858.
In the 1840–1850s, most home-based lamps burned whale oil or other animal
fats. Historically, whale-oil prices had always fluctuated wildly, but they peaked
in the mid-1850s. By some estimates, due to the over-hunting of whales, in 1860
several species were almost extinct. Whale oil sold for an average price of US$1.77
per gallon between 1845 and 1855. In contrast, lard oil sold for about US$0.90 per
gallon. Lard oil was more abundant, but it burned with a smoky, smelly flame. Michael
Dietz invented a flat-wick kerosene lamp in 1857. The Dietz lamp was arguably the
most successful of several devices designed to burn oils other than animal fats [19].
Ignacy Łukasiewicz independently developed practical kerosene lamp. By 1858–59,
Łukasiewicz lamps were replacing other forms of illumination in Austrian railway
stations.
The availability of kerosene got a sudden boost on August 27, 1859, when Edwin
L. Drake struck oil with the well he was drilling near Titusville, Pennsylvania. By
today’s standards, the well was shallow—about 69 feet (21 meters) deep and it
produced only 35 barrels per day. Drake was able to sell the oil for US$20 per barrel,
a little less than the price of lard oil and 70% less than the price of whale oil.
Drake’s oil well was not the first—according to one source, the Chinese beat
Drake by thousands of years—but it may have been the first for which the goal was
oil production, and it may have been the first well of any kind drilled through rock
with a steam powered rotary engine. In any event, the Drake well certainly triggered
the Pennsylvania oil rush. Figure 1.2 shows some of the closely spaced wells that
sprang up in 1859 in the Pioneer Run oil field a few miles from Titusville.
According to a report issued in 1860 by David Dale Owens, the state geologist
of Arkansas: “On Oil Creek in the vicinity of Titusville, Pennsylvania, oil flows out
from some wells at the rate of 75–100 gallons in 24 hours already fit for the market.
At least 2000 wells are now in progress and 200 of these are already pumping oil or
have found it.”
According to The Prize, a prize-winning book by Daniel Yergin: “When oil first
started flowing out of the wells in western Pennsylvania in the 1860s, desperate oil
men ransacked farmhouses, barns, cellars, stores, and trash yards for any kind of
barrel—molasses, beer, whiskey, cider, turpentine, sale, fish, and whatever else was
12 1 Characteristics and Historical Events
Fig. 1.2 Pioneer Run oil field in 1859. Photo used with permission from the Pennsylvania Historical
Collection and Museum Commission, Drake Well Museum Collection, Titusville, PA
handy. But as coopers began to make barrels especially for the oil trade, one standard
size emerged, and that size continues to be the norm to the present. It is 42 gallons.”
The United States has produced about 3 billion barrels since the Drake Well. In
1870, America was the world’s leading oil producer, and oil was America’s 2nd
biggest export. The first cargo of oil was exported from American in 1861, and
by 1870 Russia and the United State were the two leading countries for petroleum
development. In the 1930s, the United States was the leading producer of oil, but was
also a major consumer and therefore not a major exporter. Oil companies began to
expand their exploration interests into countries in the Middle East, Africa, Europe,
and to Canada.
In 1879, Thomas Edison invented the electric light bulb, which slowly but surely
began to replace kerosene as an illuminant. In the late 19th Century and the early
20th Century, the world’s navies began to switch from coal to fuel oil. The Anglo-
Persian Oil Company (later part of BP) was established to provide fuel for the British
navy. In 1889, Gottlieb Daimler, William Mayback and (separately) Karl Benz built
the first gasoline-powered automobiles, creating a niche market for naphtha. Inter-
nal combustion engines grew in popularity, but were too expensive for most people.
Henry Ford changed that. In 1908, the Ford Motor Company began selling Model T
automobiles for $950 each. By 1910, 50,000 cars filled the roads of America and the
growing demand for fuels pushed refining to its limits. Refining at this moment was
still primitive up until the introduction of cracking. Cracking was discovered when
engineers realized heavier fractions of crude oil could be cooked until they cracked
1.3 Use of Petroleum: A History 13
into lighter components. In the 1920s, the creation of Prohibition Act allowed for
experts in alcohol distillation to find jobs elsewhere, namely, the petroleum indus-
try. Bringing knowledge previous learned in the spirits industry, these technologists
revolutionized the petroleum field with a multitude of developments and enthusiasm
for research. Petroleum refining was well established at this point.
John D. Rockefeller and his partners started concentrating on oil refining, instead
of drilling for oil, in 1867. His company grew by taking over other local refineries
in Cleveland, Ohio and established the Standard Oil Company in 1870. Through
horizontal integration his wealth soared as kerosene and gasoline grew in importance,
and he became the richest person in the country. At the peak, Standard Oil controlled
90% of all oil in the United States. In 1911, citing violations of federal anti-trust laws
by Supreme Court, the U.S. Justice Department ordered the breakup of Rockefeller’s
Standard Oil into 34 entities, including five regional “majors”, shown in Fig. 1.3.
Among these, Standard Oil of New York (SOCONY then Mobil) and Standard Oil
of New Jersey (Esso then Exxon) were merged into ExxonMobil in 1999. Standard
Oil of Ohio (Sohio) and Standard Oil of Indiana (Amoco) were acquired by BP in the
1990s. Standard Oil of California (SOCAL) changed the name to Chevron, which
merged with Gulf Oil in 1985 and acquired Texaco in 2001. In the 1920s, the seven
leading global companies (also known as seven sisters) shown in Fig. 1.4 dominated
the oil industry worldwide. Three of the 5 largest oil companies are based in the
U.S. and are spinoffs from Standard Oil; the remaining two are in Europe. Table 1.4
lists the historical events of discovery, recovery, refining and usage of crude oil from
prehistorical era to 1911.
Organization of Petroleum Exporting Countries (OPEC) was founded in Bagdad,
Iraq by Iran, Iraq, Kuwait, Saudi Arabia and Venezuela, with headquarters in Vienna,
Austria. In addition to the five founding members, there are 10 other members cur-
rently: Algeria, Angola, Ecuador, Equatorial Guinea, Gabon, Libya, Nigeria, Qatar,
the Republic of Congo and United Arab Emirates. Indonesia suspended OPEC mem-
bership in 2016 for not agreeing with oil production cut. As of 2018, the 15 countries
accounted for 44% of global production and 81.5% of the world’s “proven” reserves.
Two-third of OPEC’s oil production and reserves are in the six Middle-Eastern coun-
Oil in 1911
Standard of New
Jersey
Standard of
Indiana
Standard of Ohio
Standard of
California
14 1 Characteristics and Historical Events
Table 1.4 Historical events of the discovery, recovery, refining and usage of crude oil
Date Description
≈ 70,000 BC According to archeologists, bitumen was used for hafting
≈ 5500 BC spears in Umm el Tlel, in present-day Syria
3200 BC Baskets were water-proofed with bitumen before 5500 BC
in Mehrgarh, in present-day Pakistan
Written descriptions of bitumen use correspond to the
advent of writing. Sumerian writing describe the use of
asphalt as an adhesive for making mosaics, for lining
water canals, sealing boats, and build roads. Egyptian
writings describe the use of pitch to grease chariot wheels,
and asphalt to embalm mummies
1500 BC In China, natural gas is used as a light and heating source
600BC Confucius writes about using bamboo poles to build
pipelines and drill 100-foot natural gas wells
347 AD Oil was produced from bamboo-drilled wells in China
c. 672 AD Byzantine Greeks invented formidable “Greek fire” by
mixing petroleum products with quicklime, which was
used to defend Constantinople from the Arab attack
1200–1300 AD The Persians mine seep oil near Baku (now in Azerbaijan)
1500–1600 AD Seep oil from the Carpathian Mountains is used in Polish
street lamps. The Chinese dig oil wells more than 2000
feet (600 meters) deep
1735 AD Oil is extracted from oil sands in Alsace, France
Early 1800s Oil is produced in United States from brine wells in
Pennsylvania
1847 James Oakes builds a “rock oil” refinery in Jacksdale,
England. The unit processes 300 gallons per day to make
“paraffin oil” for lamps. James Young builds a coal-oil
refinery in Whitburn, Scotland
1848 F.N. Semyenov drills the first “modern” oil well near Baku
1849 Canadian geologist Abraham Gesner distills kerosene
from crude oil
1854 Ignacy Lukasiewicz drills oil wells up to 150 feet (50
meters) deep at Bóbrka, Poland
1857 Michael Dietz invents a flat-wick kerosene lamp (Patent
issued in 1859)
1858 Ignacy Lukasiewicz builds a crude oil distillery in
Ulaszowice, Poland. The first oil well in North America is
drilled near Petrolia, Ontario, Canada
(continued)
1.3 Use of Petroleum: A History 15
tries that surround the oil-rich Persian Gulf. The OPEC cartel offset the dominance
of “seven sisters.” It especially influences crude oil prices.
Leading Global
Industries in the
1920s
US based Non US
companies companies
Today, about 90% of transportation fuel needs are met by oil. While petroleum
accounted for 33% of worldwide energy consumption in 2015, it was responsible
for only 3.9% of worldwide electricity generation [1]. Figure 1.5 shows the world’s
energy consumption since 1820 [20].
Petroleum’s worth as a portable, energy-dense fuel powering the vast majority of
vehicles and as the base of many industrial chemicals makes it one of the world’s
most important commodities. The condition of the oil industry depends on several
key parameters, such as overall supply and demand, the number of vehicles in the
world and the kinds of fuel they burn, net energy gain (economically useful energy
provided minus energy consumed), the political stability of oil exporting nations,
and the ability to defend oil supply lines.
The top three oil producing countries are the United States, Russia (Eurasia), and
Saudi Arabia. Figure 1.6 shows the crude oil productions of these three countries in
2010–2014. The U.S. surpassed Russia and Saudi Arabia as world’s largest producing
country in 2014, after technology advances in hydraulic fracturing to produce large
quantities of gas and oil from shale.
About 48% of the world’s readily accessible reserves are located in the Middle
East, with 46% coming from the Arab Five: Saudi Arabia, UAE, Iraq, Iran and
Kuwait. As mention above, a large portion of the world’s total oil exists as bitumen
in Canada and extra heavy oil in Venezuela.
Fig. 1.6 Crude Oil Production of United States, Saudi Arabia and Eurasia (Russia and the Former
Soviet Union) in 2010–2014
References
13. Cârciumaru M, Ion R-M, Niţu E-C, Ştefănescu R (2012) New evidence of adhesive as hafting
material on middle and upper palaeolithic artefacts from Gura Cheii-Râşnov Cave (Romania).
J Archaeol Sci. https://doi.org/10.1016/j.jas.2012.02.016
14. McIntosh JR. The ancient indus valley. New Perspectives (Understanding Ancient Civiliza-
tions) 1st Edition, 2008, ABC-CLIO, Inc. Santa Barbara, California, 57
15. Asphalt: https://en.wikipedia.org/wiki/Bitumen, Retrieved 23 Sept 2015
16. Abraham H (2015) Asphalts and allied substances: their occurrence, modes of production, uses
in the arts, and methods of testing (4th ed.). 1938. Van Nostrand Co., New York. Viewed via
https://archive.org/details/asphaltsandallie031010mbp. Retrieved 23 Sept 2015
17. Taylor FS. The origins of greek alchemy, 1937, Ambix 1, 40
18. Wikipedia: Greek fire: https://en.wikipedia.org/wiki/Greek_fire. Retrieved 31 Aug 2015
19. Pees ST. (2004) Whale Oil Versus the Others. Oil History, Samuel T. Pees, Meadville, Penn-
sylvania
20. Tverberg G. Our finite world: world energy consumption by source since 1820. http://
ourfiniteworld.com/2012/03/12/world-energy-consumption-since-1820-in-charts/. Retrieved
4 Jan 2015
Chapter 2
Crude Assay and Physical Properties
Around the Earth, different living organisms perished and drifted to the bottom of soil
or bodies of water, along with other inorganic materials. Over hundreds of millions of
years, under the influence of high pressure and temperature, some carbon-containing
portions of the dead organisms escaped full oxidation and ended up as organic matter
in sedimentary rocks, as kerogen, oil, or gas. Due to differing climates, organisms, and
depositional environments, the compositions of oil depend drastically with region
and can vary very significantly. Hence, it is important to determine physical and
chemical characteristics of crude oil through a crude oil assay (crude assay).
There are hundreds of different crude oils produced in the world with large variations
in properties and composition. Crudes typically are named for their source country
and reservoir. Both physical and chemical properties of crude oils vary with location,
depth, and the age of the oil field.
Characteristics of crude oil that can indicate these differences include: boiling
point, calorific value (heating value), API gravity (density), pour point, cloud point,
freeze point, aniline point, smoke point, flash point, viscosity, color and fluorescence,
Reid vapor pressure, refractive index, sulfur content, nitrogen content, metal content,
salt content, micro carbon residue, optical activities, and total acid number [1]. These
crude oil bulk properties measured by testing laboratories, along with distillation and
product fractionation data at pilot plants, are compiled into a crude oil assay (or crude
assay) to characterize a specific crude oil.
A crude assay can be an inspection assay or a comprehensive assay (full assay).
Testing can include characterization of the whole crude oil and/or its various boiling
fractions produced from fractional distillation. High quality assay data are important
to refinery planners and oil traders, who select crudes by comparing their properties
with refinery specifications and constraints as well as meeting environmental and
other standards. Sulfur content and total acid number (TAN) are common constraints,
© Springer Nature Switzerland AG 2019 19
C. S. Hsu and P. R. Robinson, Petroleum Science and Technology,
https://doi.org/10.1007/978-3-030-16275-7_2
20 2 Crude Assay and Physical Properties
because they relate to corrosion in piping and process equipment. The refiners make
changes in plant operations based on the crude assay data, to meet process and
product requirements. Hence, crude assays provide the basis upon which companies
and traders negotiate contracts, which include pricing and possible penalties due to
impurities and other undesired properties. However, traders might purchase off-spec
crudes for subsequent trading.
The assays determine selected chemical and physical properties of whole crudes
and several distilled fractions. The fractions correspond to boiling ranges for com-
mon fuels. Full assays are extensive, and especially important for new crudes. For
inspection assays, just a few tests are conducted. In a whole crude assay, analyses
are done by combining atmospheric and vacuum distillation runs to provide a true
boiling-point (TBP) distillation data. The extent of an assay depends on customer
need and affordability. At minimum, the assay should contain a distillation curve,
typically a TBP curve, and a specific gravity curve. The common assay inspections
are discussed below.
2.2 Distillation
Distillation is the primary method of separating crude oil into useful products. It
utilizes differences in boiling points (volatility) of fractions. Boiling point distribu-
tion of the components, determined by fractional distillation, is the most important
characteristic of petroleum. Large scale and continuous distillation is also the most
basic process in any refinery to separate crude oil into fractions of different boiling
ranges, shown in Fig. 2.1. Table 2.1 lists the major distillation products (fractions)
with their boiling ranges.
Naphtha of high octane-number is a main component of gasoline. It is composed of
hydrocarbons mainly from C5 to C10 , with some C11 and C12 . The carbon numbers of
heavy naphtha extend to mid-teens, overlapping with those of kerosene. Kerosene is
a middle distillate fraction of crude oil with C10 to C16 . It is mainly used for heating
oil, jet fuel, and diesel. It is the first fraction to show an appreciable increase in
the cyclic hydrocarbons that dominate the heavier fractions. Aromatics in kerosene
range from 10 to 40%. Light gas oils, C14 to C18 , are used in both jet fuels and
diesel fuels. The C20 to C50 range contains diesel fuels, heating oils, lubricating oils,
paraffin waxes and some asphalts. Residuum (or resid) includes resins, asphaltenes
and waxes, and is the most complex and least understood fraction of petroleum. The
wax fraction in most residua is about half that in the lube oil fraction. Asphaltenes are
dark brown to black amorphous solids. The resins may be light to dark colored, thick,
viscous substances to amorphous solids. The resins and the asphaltenes contain about
half of the total nitrogen and sulfur in crude oil. Heavy crude oils invariably have
more nitrogen and sulfur. Residua frequently contain over 5% oxygen. Resins are
the highest of the fractions in oxygenated compounds, while asphaltenes are highest
in sulfur compounds.
2.2 Distillation 21
The amount of valuable products, or the yields of fuels and lubricant oils that can
be produced from a crude oil simply by distillation is a major factor in determining
the value of the oil.
Smaller scale or laboratory scale distillation methods are described in the Annual
Book of ASTM Standards [2]. ASTM stands for the American Society for Testing
Materials, now called ASTM International. The methods include ASTM D2892
for atmospheric distillation and ASTM D5236 for vacuum distillation. The ASTM
D2892 (15/5 distillation) method uses a 15 theoretical plate fractionation column
with a 5:1 reflux ratio. It determines accurate boiling points and yields of distillation
fractions. The highest temperature is limited to ~350 °C (650 °F) to avoid thermal
22 2 Crude Assay and Physical Properties
Fig. 2.3 Simulated distillation by gas chromatography (the numbers on top of the peaks are carbon
numbers of normal paraffins)
The following lists some of the most important tests for physical properties of crude
oil or its fractions.
• Specific gravity and API gravity
Specific gravity, or relative density, is the ratio of the mass of a given volume of the
oil to the mass of an equal volume of water at a specific temperature, which is often
60 °F. A commonly used tool for measuring specific gravity includes a temperature
bath heated to the desired temperature, a metal tube to fill with the test sample, and a
hydrometer which indicates the specific gravity. This test determines specific gravity
which is then converted to API gravity, a measurement that has become a standard
in the petroleum industry.
API gravity is another notation of the density of the oil, which was established
by the American Petroleum Institute (API) as a measure of how heavy or light a
petroleum liquid is compared to water. It is defined by ASTM D287/1298 as
◦
API = (141.5/specific gravity at 60◦ F) − 131.5
The specific gravity of water at 60 °F is 1.0; hence, its °API is 10. The °API
of crude oils ranges from <10 for asphaltic crude to >50°API for condensate. Most
crudes are in the 20–45°API range. Condensates range between 50 and 70°API. The
API gravity of heavy oils falls in the range of 10–15° or <20°API, and for bitumen it
falls in the range of 5–10°. In 2010, the World Energy Council defined “extra heavy
oil” as crude oil having a gravity less than 10° API and reservoir viscosity no more
than 10,000 cP, with a lower API limit of 4°, so it sinks in water rather than floating
on it. API gravity is designed so that its value is more or less proportional to the
commercial value of the crude oil. A denser oil has a lower °API, and hence, the
commercial value of the oil is lower.
• Viscosity
Viscosity is a measurement of fluid resistance to shearing flow using a small cap-
illary tube (viscometer) for the sample to flow through. This measures the kinematic
viscosity, that is, the ratio of dynamic (shear) viscosity to the density of the fluid, in
centistoke (cSt) at a given temperature. It is a very important oil property.
In oil production, viscosity determines the flow of oil and gas through the reservoir,
thus, the amount and rate at which oil and gas can be produced.
In engines, viscosity determines how easily the oil is pumped to the working
components, how easily it passes through filters, and how quickly it drains back to the
engine. The lower the viscosity, the easier all this will happen. That is why cold starts
are so critical to an engine: because the oil is cold and so relatively thick that it loses
its lubricity. A fluid’s viscosity is directly related to its load-carrying capabilities. The
greater the viscosity, the greater the loads it can withstand. The viscosity must be
adequate to separate moving parts under normal operating conditions (temperature
26 2 Crude Assay and Physical Properties
Fig. 2.5 Viscosity and density of crude oils. cp: centipoise [4]
and speed). Knowing that a fluid’s viscosity is directly related to its ability to carry
a load, one would think that the more viscous a fluid, the better it is. The fact is,
the use of a high-viscosity fluid can be just as detrimental as using too light an oil.
Viscosity index (VI), that is, the change of viscosity with temperature, is a critical
property of lubricant base oil. The higher the VI, the lower is the viscosity change
with temperature. High performance and synthetic lube oils have high VI’s.
For the same homologous series of hydrocarbons, the greater the molecular weight
of the compound, the greater the viscosity. When the molecular weight is similar,
the viscosity of a cyclic molecule is larger than that of the chain-like molecule.
The greater the number of rings is, the greater the viscosity is. Therefore, the “ring
structure is the viscosity carrier”.
Viscosity is a parameter of oil mobility, also an indispensable physical property
for crude oil recovery and refining processes. The viscosity of the oil decreases as
its temperature increases.
The viscosity and density of crude oils are closed related, as shown in Fig. 2.5.
Oil sand (tar sand) bitumen is the most viscous and has the highest density (lowest
API gravity) . On the other hand, conventional crude oils have less viscosity with
lower density (or API gravity) . Oil sand was defined by the United States Congress
in 1976 as rock types that contain extremely viscous hydrocarbons which cannot be
recovered by conventional oil production methods, including enhanced recovery. Oil
sand bitumen and extra heavy crude oils are therefore referred to as unconventional
oils.
Viscosity for conventional crude oils ranges from 10–100 mPa (884–934 kg/m3 ),
and for heavy crude oils from 1000 to 10,000 mPa (966–1000 kg/m3 ). Tar sand
bitumen have viscosities that ranges from 100,000 to 1 million mPa and density over
1000 kg/m3 .
• Total Sulfur (atom) content
The sulfur (atom) content of crude oils is in the range of 0.1–5.0 wt%, measured
by x-ray fluorescence (ASTM D4294 or D5291). The terms “sweet” and “sour” are
historical terms which refer to the taste of crude oil as a function its sulfur content.
Indeed, early prospectors would taste oil to determine its quality. Low sulfur oil
actually tasted sweet. Crude oil is currently considered sweet if it contains less than
0.5% sulfur.
Sulfur compounds can affect the activity of catalysts. As mentioned earlier, com-
bustion converts organic sulfur compounds into sulfur oxides (SOx ), which can react
2.3 Physical Testing 27
with moisture (H2 O) to form fine sulfuric acid and sulfate particulates (aerosols) in
air. The particulates are transported hundreds of miles, eventually returning to the
earth as wet or dry acid deposition. The deposition, also known as acid rain, harms
buildings, trees and other plants, fish, and land animals. Similarly, organic nitrogen
compounds are converted by combustion into nitrogen oxides (NOx ), which reacts
with volatile organic compounds (VOC) by photochemical reactions to form smog.
Both SOx and NOx form particulate matter (PM) in air. PM with diameters smaller
than 2.5 μm (PM 2.5) are inhalable particles that can reach bronchial tubes to damage
lungs, causing respiratory ailments.
Sweet crude is easier to refine and safer to extract and transport than sour crude.
Because sulfur is corrosive, light crude also causes less damage to refineries and thus
results in lower maintenance costs over time.
Major locations where sweet crude is found include the Appalachian Basin in
Eastern North America, Western Texas, the Bakken Formation of North Dakota and
Saskatchewan, the North Sea of Europe, North Africa, Australia, and the Far East,
including Malaysia and Indonesia. African crudes tend to be relatively sweet: Bonny
Light, the main Nigerian crude, contains about 0.16 wt% sulfur. Brega, the main
Libya crude, contains about 0.2 wt% sulfur.
Sour crude oils have more than 0.5% total organic sulfur not including dissolved
hydrogen sulfide. Sour crude also contains more carbon dioxide. Most sulfur in
crude oil is actually bonded to carbon atoms as sulfur-containing hydrocarbons.
Nevertheless, high quantities of hydrogen sulfide in sour crude can pose serious
health problems or even be fatal.
• Hydrogen sulfide
Hydrogen sulfide is famous for its “rotten egg” smell, which is only noticed at low
concentrations. At moderate concentrations, hydrogen sulfide can cause respiratory
and nerve damage. At high concentrations, it is instantly fatal. Hydrogen sulfide is
so much of a risk that sour crude has to be stabilized via removal of hydrogen sulfide
before it can be transported by pipelines and oil tankers. Sour crude is more common
in the Gulf of Mexico, Mexico, South America, and Canada. Middle Eastern crudes
tend to be relatively sour.
• Mercaptans
Mercaptans (thiols) smell like as garlic or rotten eggs and can be toxic. A trace
amount can be used as an odorant of natural gas for detection; natural gas is odorless in
pure form. Mercaptan sulfur is measured by potentiometric titration of an isopropanol
solution of a hydrocarbon sample containing a small amount of NH4 OH. The solution
is then titrated with a silver nitrate solution. Mercaptans are removed from sour
crudes and oxidized into sulfides through the Merox process or into elemental sulfur
by LOCAT® and other methods, discussed in Sect. 7 of Chapter 9, to reduce odor
prior to transportation.
28 2 Crude Assay and Physical Properties
• Nitrogen content
Nitrogen-containing compounds can cause poisoning of acidic catalysts due to
their basicity. Nitrogen content is normally determined by oxidative combustion and
chemiluminescence detection (ASTM D3228 or D4629).
• Total Acid Number (TAN)
Total Acid Number (TAN) is the amount of KOH in mg that is needed to neutralize
the acid in a gram of oil dissolved in toluene/isopropanol/water (ASTM D664).
Typically, values are 0.05–6.0 mg KOH per gram of the sample. High TAN crudes
are purchased and processed carefully due to possible corrosion problems.
• Metal Content
Metal Content ranges from a few to several hundred ppm. It is measured by induc-
tively coupled plasma atomic emission spectroscopy (ICP-AES) (ASTM D5708) or
x-ray fluorescence (ASTM D5863). Nickel and vanadium are common in crude oils;
they can severely affect catalyst activity.
• Salt Content
Salt Content is measured by conductivity of a crude oil sample dissolved in water
compared to reference salt solutions (ASTM D3230) to determine crude oil corro-
sivity that can lead to shorter life times of pipes and pumps in the refinery. Desalting
is needed when the salt content is greater than 30 ppm to bring it down to 2 ppm.
• Micro Carbon Residue (MCR)
Micro Carbon Residue (MCR) is measured by the Conradson carbon (CCR or
ConCarbon) method ASTM D189.
MCR is a measurement of hydrocarbon mixtures’ tendency to leave carbon
deposits (coke) when burned as fuel or subjected to intense heat in a processing
unit such as a catalytic cracker. The ConCarbon test involves destructive distilla-
tion, i.e., subjection to high temperature, which causes cracking, coking, and drives
off any volatile hydrocarbons produced, and weighing the residue which remains.
A somewhat similar test, Ramsbottom carbon, also measures mixtures tendency to
form coke.
• Calorific Value (Heating Value)
In the fuel fractions of petroleum, the heating value is of importance. Paraffins
have higher calorific values than aromatics. The average value for a crude oil is
2100–2230 kcal/kg, compared with 1170–1670 kcal/kg of bituminous coal.
• Pour Point
The pour point of a liquid is the temperature at which it becomes semi solid and
loses its fluidity. It is the lowest temperature at which the oil no longer moves. The
pour point is determined as 3° above the point at which a sample no longer moves
when inverted (ASTM D97). It is important for pipeline transportation from source
2.3 Physical Testing 29
to loading ports. In crude oil a high pour point is generally associated with a high
paraffin content, typically found in crude derived from a larger proportion of plant
material as in Type III kerogen.
• Cloud Point
In the petroleum industry, cloud point refers to the temperature below which wax in
a sample, particularly jet fuel and diesel, forms a cloudy appearance. The presence
of solidified waxes thickens the oil and clogs fuel filters and injectors in engines.
The wax also accumulates on cold surfaces (e.g. causing pipeline or heat exchanger
fouling) and forms an emulsion with water. Therefore, cloud point indicates the
tendency of the oil to plug filters or small orifices at cold operating temperatures.
In crude or heavy oils, cloud point is synonymous with wax appearance temper-
ature (WAT) and wax precipitation temperature (WPT). It is determined by ASTM
D2500 or D5773 method as the temperature at which a haze appears in a sample due
to formation of wax crystals by cooling.
• Freeze Point
Freeze Point is the temperature at which crystals start to form in hydrocarbon
liquid and then disappear when heated (ASTM D2386).
• Aniline Point
Aniline Point is the lowest temperature at which aniline and an oil are completely
miscible. It is measured by ASTM D611. The mixture is heated and stirred until
homogeneous, then it is cooled with stirring until the two liquids separate. The
temperature at which such separation occurs is the aniline point. For clear samples,
the aniline point is that at which the mixture suddenly becomes cloudy. In darker
samples, the apparatus becomes more complicated, as it is less obvious when the
mixture is cloudy. It is used in some specifications as an indication of aromatic
content. The lower the aniline point, the greater is the aromatic content.
• Smoke Point
Smoke Point is performed on jet fuels and kerosene cuts to determine clean burning
by measuring flame height with a standard wick, expressed in mm, in a lamp without
smoke forming (ASTM D1322). To test for smoke point, a sample is burned in a
lamp that is precisely calibrated using hydrocarbons with known smoke points. The
maximum height of the flame without production of smoke is recorded to the nearest
0.5 mm. This is important in determining the composition of crude oil. Samples with
lower smoke points are more aromatic. Smoke point is usually most important in jet
fuel. Smoke can shorten the lifetime of engine parts, therefore higher smoke points
are favorable.
• Flash Point
Flash point is the temperature at which combustion occurs when the hydrocarbon
mixture is exposed to air or oxygen.
30 2 Crude Assay and Physical Properties
altitudes increase vaporization, which can lead to “vapor lock.” While handling crude
oil, vapor pressure is of high importance for safety reasons.
Note that RVP differs slightly from true vapor pressure (TVP) of a liquid. TVP is
measured in the presence of water vapor and air during sample vaporization in the
confined space of the test equipment. Hence, the RVP is the absolute vapor pressure
and TVP is the partial vapor pressure.
Table 2.2 exhibits an example assay report template; there is no standard assay testing
grid—each refinery or trading company has its own. Note that not all the tests need
to be done for all samples. The tests are requested and performed depending on the
needed information for processes and products.
Figure 2.6 shows an example of crude assay report of a benchmark crude, Brent
blend, and its fractions [5]. The yields of different fractions (cuts) is expressed by
volume %. It can be seen that paraffins are concentrated in the lowest distillation cuts
as “paraffinic”, with increasing naphthene and aromatic contents in higher distillation
cuts as “asphaltic”.
Table 2.3 presents selected bulk physical and chemical properties for 21 crude oils.
Athabasca is a heavy oil, with a specific gravity >1.0. In the table, sulfur contents
range from 0.14 to 5.3 wt%, and nitrogen contents range from nil to 0.81 wt%.
Specific gravities range from 0.798 to 1.014. In sweet crudes such as Tapis, the
sulfur content is low. Sour crudes have more sulfur, which gives them a tart taste; in
the old days, prospectors did indeed characterize crude oil by tasting it. The crude
oils in China (Henan, Liaohe, Shengli, Xinjiang, etc.) have relatively higher nitrogen
contents in general compared to the crude oils outside China. The high nitrogen
(relative to sulfur) presents operational challenges in refining, where it is converted
into ammonia. The usual way to remove ammonia is in the form of ammonium
bisulfide (NH4 SH). The NH4 SH dissolves in wash water and is transported to sulfur
plants. If there is more NH3 than H2 S, the NH3 is not completely removed in this
fashion and remains in process gas streams, where it can accelerate corrosion.
Table 2.2 Example crude assay report template
32
Whole crude Light naphtha Medium naphtha Heavy naphtha Kero AGO LVGO HVGO VR AR
True Boiling Initial 10 80 150 200 260 340 450 570 340
Point, °C
True Boiling Final 80 150 200 260 340 450 570 End End
Point, °C
True Boiling Initial 55 175 300 400 500 650 850 1050 650
Point, °F
True Boiling Final 175 300 400 500 650 850 1050
Point, °F
Yield of Cut (wt% x x x x x x x x x
of Crude)
Yield of Cut x x x x x x x x x
(vol.% of Crude)
Gravity, °API x x x x x x x x x x
Specific Gravity x x x x x x x x x x
Sulfur, wt% x x x x x x x x x x
Nitrogen, ppm x x x x x x x x x
Viscosity @ 50°C x x x x x x x x
(122°F), cSt
Viscosity @ x x x x x x x x
135°C (275°F),
cSt
Freeze Point, °C x x x x
Freeze Point, °F x x x x
Pour Point, °C x x x x x x x x
Pour Point, °F x x x x x x x x
2 Crude Assay and Physical Properties
(continued)
Table 2.2 (continued)
Whole crude Light naphtha Medium naphtha Heavy naphtha Kero AGO LVGO HVGO VR AR
Smoke Point, mm x x x
Aniline Point, °C x x x x x x
Aniline Point, °F x x x x x x
Cetane Index, x x x
ASTM D976
Diesel Index x x x x x x
Characterization x x x x x x x x x x
2.5 Bulk Properties of Crude Oil
Factor (K)
Research Octane x x x
Number, Clear
Motor Octane x x
Number, Clear
Paraffins, vol.% x x x x x x
Naphthenes, x x x x x x x
vol.%
Aromatics, vol.% x x x x x x x
Heptane x x x
Asphaltenes, wt%
Micro Carbon x x x
Residue, wt%
Ramsbottom x x x
Carbon, wt%
Vanadium, ppm x x x
Nickel, ppm x x x
Iron, ppm x x x
33
34 2 Crude Assay and Physical Properties
Fig. 2.6 Crude assay results of a Louisiana crude oil and its fractions [7]
Among these crude oils, West Texas Intermediate (WTI) and Brent from North
Sea are commonly used in the commercial or merchandise communities around
the globe as benchmark crudes for the crude oil prices which change constantly
depending on supply and demand. Both benchmark crudes have relatively low sulfur
and nitrogen compared to other oils. The price of a specific crude oil can be higher
or lower than these reference crudes. Value is set primarily by the results of crude
assays. In general, the crude oils with low API gravity, high sulfur content and high
acid content are sold at deep discounts compared to the reference oils.
2.5 Bulk Properties of Crude Oil 35
The weight percent data of sulfur and nitrogen contents in the crude oils listed
in Table 2.3 can also be presented in graphical form as in Fig. 2.7. Both sulfur
and nitrogen correlate inversely with API gravity, but for this particular dataset, the
correlations are rough due to wide scattering of data points, especially for sulfur.
Table 2.4 exhibits one example of reporting selected assay data for the whole
crude and its different residue cuts in a Mexican crude oil. As expected, API gravity
decreases with cut point; that is, the deeper cut point, the lower the API gravity.
The API of the resid obtained at cut point at 650 °F (lowest cut point) is 17.3
while at 1050 °F (highest cut point) it is 7.1. Sulfur content, nitrogen content, metal
content, viscosity, pour point and Conradson carbon (ConCarbon) residue increase
with increasing boiling point.
36 2 Crude Assay and Physical Properties
Fig. 2.7 Sulfur and nitrogen versus API gravity for selected crude oils. The top correlation line is
for nitrogen and the bottom one is for sulfur, which is more scattered (has a lower R2 value)
Table 2.4 Properties of a Mexican crude oil and its residua at different cut points [5]
Whole Crude Residua
650 °F 950 °F 1050 °F
Yield, vol.% 100.0 48.9 23.8 17.9
Sulfur. wt% 1.08 1.78 2.35 2.59
Nitrogen, wt% 0.33 0.52 0.60
API gravity 31.6 17.3 9.9 7.1
Conradson Carbon wt% 9.3 17.2 21.6
Vanadium, ppm 185 450
Nickel, ppm 25 64
Kinematic Viscosity at 100 °F 10.2 890
Kinematic Viscosity at 210 °F 35 1010 7959
Pour Point, °F −5 45 95 120
References 37
References
1. Hsu CS (1995) Hydrocarbons. In: Encyclopedia of analytical science, premiere edition, Aca-
demic Press, London, pp 2028–2034
2. Annual Book of ASTM Standards, section 5: Petroleum products, liquid fuels, and lubricants,
Vol. 05.01 and Vol. 05.02. In: American society for testing and materials (ASTM) International,
West Conshohocken, PA, 2016
3. Gray JH, Handwerk GE, Kaiser MJ (2007) Petroleum refining—technology and economics, 5th
Edition, CRC Press
4. Lopes MS, Savioli Lopes M, Maciel Filho R, Wolf Maciel MR, Median LC (2012) Extension
of the TBP curve of petroleum using the correlation DESTMOL. Procidea Eng 42:726–732
5. Speight JG (2006) The chemistry and technology of petroleum, 4th edn. Marcel Dekker, Revised
and Expanded
6. Leffler WL (2008) Petroleum refining in nontechnical language. Fourth Edition, PennWell
7. Assays available for download. https://corporate.exxonmobil.com/en/company/worldwide-
operations/crude-oils/assays. (Retrieved 20 Sept 2018)
Chapter 3
Chemical Composition
Table 3.1 Overall elemental Element All petroleum Tar sand bitumen
composition by weight
percent [3] Carbon 83–87% 83.4 ± 0.5%
Hydrogen 10–14% 10.4 ± 0.2%
Sulfur 0.05–6.0% 5.0 ± 0.5%
Nitrogen 0.1–2.0% 0.4 ± 0.2%
Oxygena 0.05–1.5% 1.0 ± 0.2%
Metals (Ni and V) <1000 ppm >1000 ppm
a Oxygen is calculated by difference. Caution is needed in sample
handling, especially for prolong exposure to air during storage
Table 3.1 shows the average overall elemental composition of all crude oils and tar
sand bitumen from Canada. It is noticeable the sulfur, nitrogen, oxygen and metal
contents in tar sand bitumen are significantly higher than the average of all petroleum.
More than any other element, carbon binds to itself to form straight chains,
branched chains, rings, and complex three-dimensional structures. The most complex
molecules are biological—proteins, carbohydrates (including cellulose and hemi-
cellulose), lipids, nucleic acids, and lignin. This is significant, because petroleum
was formed from the remains of ancient microorganisms—primarily planktons and
algae. However, petroleum is derived from these biomolecules after losing func-
tional groups, leaving behind carbon skeletons that serve as biological markers or
biomarkers. The remnants of cellulose, hemicellulose and lignin can be found in
another hydrocarbon resource—coal.
Since hydrogen is a much lighter element than the others, oils with higher hydro-
gen content have lower specific gravities, i.e., higher API gravities. High API gravity
oils have a high naphtha and kerosene/middle distillate (light ends) content and low
residuum (heavy ends) content whereas low API gravity oils are low in naphtha and
kerosene/middle distillate content and high in residuum.
Distillation is the first step in characterizing petroleum. Distillation separates
petroleum into different boiling fractions: gas, naphtha, kerosene/jet fuel, light
gas oil/diesel, heavy gas oil, lubricating oil, and residuum. The physical and chem-
ical properties of petroleum vary with location, depth, and age of an oil field. The
differences in these properties can be attributed to the distributions of the different
sizes and types of hydrocarbons.
Petroleum molecules can be classified in several ways. They can be divided into
saturated, unsaturated, and polar compounds. They can also be classified as paraffins,
olefins, naphthenes, aromatics, polynaphthenes, polyaromatics, naphthenoaromat-
ics, and heteroatom-containing compounds. Saturated hydrocarbons can be acyclic
3.2 Molecular Composition 41
Paraffins (alkanes) have a general formula of Cn H2n+2 . Figure 3.1 shows molecular
structures of some simple alkanes. The simplest paraffin is methane with a single
carbon atom. Methane is the major component of natural gas. The next member in
the alkane family is ethane with 2 carbon atoms. After that comes propane, with 3
carbon atoms. When the carbon number reaches 4, isomers are possible. Isomers are
chemical compounds with the same molecular formula but different structures.
In “normal” paraffins, no carbon atom is connected to more than two other carbon
atoms. In simple two-dimensional representations, the carbons are drawn in a straight
chain as linear structures. In isoparaffins, at least one carbon atom is connected
to three or four other carbon atoms. In two-dimensional diagrams, the structure is
branched. Carbon atoms connected to only one other carbon, such as the end-of-chain
carbons in n-paraffins, are called primary (1°). Carbon atoms connected to two other
carbons are called secondary (2°), those connected to three other carbons are called
tertiary (3°), and those connected to four other carbons are called quarternary (4°).
For example, C4 H10 includes normal butane (n-C4 ), in which all carbon atoms are
primary or secondary, and isobutane (methyl propane or i-C4 ), in which the central
carbon atom is tertiary. C5 H12 can have three isomers, normal pentane, isopentane
(2-methyl butane) and neopentane (2,2-dimethyl propane), as shown in Fig. 3.1. The
central carbon in neopentane is quarternary.
“Isooctane” in the figure is just one of several isooctanes. Its official (IUPAC) name
is actually 2,2,4-trimethylpentane. However, this particular molecule is especially
important, because it serves as a yardstick of gasoline combustion performance in
spark-ignition engines, namely, “Octane Number”. For isoparaffins, 2-methyl and 3-
methyl alkanes are most abundant. 4-methyl derivatives are present in small amounts,
if at all.
Petroleum also contains multi-methyl alkanes. Pristane (2,6,10,14-
tetramethylpentadecane) and phytane (2,6,10,14-tetramethylhexadecane) contain
isoprene skeleton structures, referred to as isoprenoid hydrocarbons. They are
important acyclic biomarkers for indicating the origin and depositional environment
of petroleum.
Different isomers have different boiling points and melting points, as illustrated
in Tables 3.2 and 3.3. Normal butane (n-butane) boils below the melting point
of ice. It is blended into gasoline in cold weather to increase vapor pressure, i.e.,
Reid vapor pressure, thereby improving the ignition behavior of spark plug internal
42 3 Chemical Composition
Table 3.4 Number of possible isomers of paraffins with a specific carbon number
Carbon number Boiling point of n-Paraffins Number of isomers
°C °F
5 36 97 3
10 174 345 75
15 271 519 4347
20 344 651 366,319
25 402 755 36,797,588
30 450 841 4,111,846,763
40 525 977 62,491,178,805,831
for octane number calculations based on molecular composition. They become less
important beyond gasoline range (>C10 ).
Arenes, or aromatics, contain one or more benzene rings. Aromatics are also unsat-
urated hydrocarbons that will react to add hydrogen or other elements to the ring.
Aromatics tend to be concentrated in heavy high boiling fractions of petroleum
such as gas oil, lubricating oil, and residuum. In naphtha fraction, aromatics have the
high octane-rating of hydrocarbon types, so they are valuable components in gasoline
blends. However, they are undesirable in the lubricating oil range because they have
the highest change in viscosity with temperature, that is, lowest viscosity index (VI),
of all the hydrocarbons. Asphaltenes are the most aromatic of the fractions. Average
oil tends to have more paraffins in the gasoline fraction and more aromatics and
asphaltics in the residuum.
The simplest aromatic molecule is benzene, containing 6 carbon atoms and 6
hydrogen atoms with 3 conjugated double bonds around the ring. The simplest
2- and 3-ring aromatics are naphthalene, anthracene, and phenanthrene. Phenan-
threnes are commonly found in crude oils while their isomer anthracenes are
more common in coal. Aromatics, like naphthenes, can have alkyl (either linear
or branched) substituents attached to the core in place of hydrogen at the substi-
tution points. Figure 3.3 presents some aromatics and naphthenes found in crude
oils.
Aromatics of greater than 2 rings can have condensed structures as shown in
Fig. 3.4. There are two kinds of condensation, peri- and cata-condensed [4]. These
condensed multiring aromatics are generally known as polynuclear aromatics (PNA),
heavy polynuclear aromatics (HPNA). or polynuclear aromatic hydrocarbons (PAH).
Many of the PAH isomers are carcinogenic, mutagenic or toxic, and are listed as
high-priority pollutants by the U.S. Environmental Protection Agency (EPA).
Figure 3.5 shows another set of ring compounds. The –R groups in the figures rep-
resent short or long alkyl chains. The long alkyl chains tend to be branched. Some
46 3 Chemical Composition
cata-condensed
peri-condensed
compounds contain both aromatic and naphthenic rings, such as indans, dinaph-
thobenzenes and dinaphthonaphthalenes. They are called naphthenoaromatics or
hydroaromatics and are formed by partial hydrogenation of aromatic hydrocarbons.
In most cases, the hydrogenation of aromatics is reversible; in some refining pro-
cesses, naphthenoaromatics come from partial dehydrogenation of polynaphthenes.
3.2 Molecular Composition 47
Some compounds contain hetero atoms, mainly sulfur, nitrogen and oxygen, as
shown in Fig. 3.6. These compounds can be aliphatic and aromatic. Trace amounts
of metal chelated compounds, most noticeably nickel and vanadium porphyrins,
have also been identified. Heteroatom-containing compounds have been called “non-
hydrocarbons” in the literature because they contain elements more than just carbon
and hydrogen. They are not straightly hydrocarbons per se.
Sulfur is the third most abundant atomic constituent of crude oils. It is mostly present
in the medium and heavy fractions of crude oils. In the low and medium molecular
ranges, sulfur is associated only with carbon and hydrogen, while in the heavier
48 3 Chemical Composition
Thiophenes are more stable than mercaptans, sulfides and disulfides. They are
removed by hydrodesulfurization. Figure 3.7 lists some of the representative sulfur
organics found in petroleum.
The total sulfur in crude oil varies from below 0.05% (by weight), as in some
Pennsylvania oils, to about 2% for average Middle Eastern crudes, and up to 5% or
more in heavy Mexican or Mississippi oils. Generally, the higher the specific gravity
of the crude oil, the greater is its sulfur content (see Fig. 1.9). The excess sulfur must
be removed from crude oil during refining, because otherwise sulfur oxides released
into the atmosphere during combustion of the oil would generate sulfur oxides, which
would turn to sulfuric or sulfurous acid with moisture and return to the earth as acid
rain (a general term for wet deposition), a major pollutant. The sulfates can also
attach to the smog particles, particularly PM 2.5), harmful to human health.
Nitrogen is present in almost all crude oils, usually in quantities of less than 0.1%
by weight. As with sulfur compounds, in general the higher the specific gravity of
the crude oil, the greater is its nitrogen content (Fig. 1.9). Except for molecular
nitrogen (N2 ), the simplest stable nitrogen-containing molecule is ammonia (NH3 ).
50 3 Chemical Composition
Amines are a class of organic compounds where the nitrogen atoms in ammonia
are replaced by alkyl groups. In primary amines, only one hydrogen atom in the
molecule is replaced by an alkyl group, linear or branched. Those in which two
and three hydrogen atoms are replaced by alkyl groups are secondary and tertiary
amines, respectively. There are amines with in which the nitrogen atom is connected
to a naphthenic ring, such as cyclohexyl amine.
Aromatic nitrogen compounds are generally divided into basic and non-basic,
as determined by titration with perchloric acid in a 50:50 glacial acetic acid and
benzene. Basic nitrogen compounds generally contain nitrogen atom with a pair of
nonbonding electrons. They are of most concern to the refiners, because many cata-
lysts used in upgrading processes contain acidic sites. Basic nitrogen compounds can
react with those sites by coordinating nonbonding electrons with empty d-orbitals on
the catalysts (Lewis acid sites) to deactivate the catalysts. Non-basic nitrogen com-
pounds containing nitrogen atoms that are bonded with hydrogen are more benign,
but can be converted into basic nitrogen compounds during processing, especially
3.2 Molecular Composition 51
The oxygen content of crude oil is usually less than 2% by weight and is present
as part of the heavier hydrocarbon compounds in most cases. For this reason, the
heavier oils contain the most oxygen.
Oxygen-containing compounds in crude oils are mainly furans, phenols and naph-
thenic acids. Other acids, such as fatty acids and aromatic acids, can also be present
in some crude oils. The presence of ketones, esters, ethers and anhydrides have been
reported in residua, but they can be generated by sample treatment or handling;
oxygen compounds can be formed from prolonged exposure to air during or after
production of crude oil.
Furans are normally present in crude oils in small amounts and are easy to remove
by hydroprocessing. Phenols are key components in coal derived from lignin in ligno-
cellulosic biomass, but insignificant in conventional crude oils. Many of the recently
discovered and recovered oils from Africa, Asia and offshore contain significant
amounts of naphthenic acids, shown in Fig. 3.9, [5] which cause corrosion concerns.
52 3 Chemical Composition
Unlike total oxygen content, which is difficult or expensive to obtain reliably, the
acid content can be much more conveniently obtained by titration.
Many metallic elements are found in crude oils, including most of those present
in seawater. This is probably due to the close association between seawater and
the organic forms from which oil is generated. Among the most common metallic
elements in oil are vanadium and nickel that show particularly deleterious effects
on catalysts, such as poisoning, excessive gas and coke formation. Crudes delivered
to refineries can contain substantial amounts of iron, but much of that comes from
corrosion during shipping and storage. Other metals in small amounts include copper,
manganese, lead, mercury, as well as nonmetals, such as arsenic and silicon.
Metals can be present in petroleum as chelates, complexes or other coordination
compounds. Although many of these organometallics remain to be characterized,
the most recognized are porphyrins, which have a tetrapyrrole structure in which the
nitrogen atoms form chelates with the metal atoms (Fig. 3.10). The most common
porphyrins contain nickel(II) and vanadium(V) (as vanadyl). Less common are por-
phyrins containing iron(II) and copper(II). Vanadium is dominant in marine-sourced
oils (Ni/V < 1) while nickel is dominant in lacustrine oils (Ni/V > 1) [6]. Metallo-
porphyrins are present in heavy petroleum fractions: vacuum gas oil and resid. They
are serious catalyst poisons.
Arsenic and mercury are present as alkyls. Both are potent catalyst poisons—five
to ten times worse than Ni, V, or Fe on a weight basis. Arsenic and mercury alkyls
are relatively volatile, so they show up not just in residue, but in lighter fractions as
well. Arsenic is common in Green River shale oil and crude oils from Canada, West
Africa, and the Ukraine. Mercury is a major problem in natural gas condensate in
Thailand.
Silicon is not a metal, strictly speaking, and it is not common in raw crude oil. But
organosilicon compounds serve as flow improvers in production facilities, pipelines,
and transportation hubs. Silicon degrades the activity and structural integrity of cer-
tain refinery catalysts. In one respect, silicon isn’t as bad as other metals, because
catalysts can tolerate more of it. Ni, V, Cu, and certain forms of iron can be rejected
by distillation, but like arsenic and mercury alkyls, silicon compounds are too light
for that. The only practical way to remove them is by chemisorption onto special
guard materials.
3.2.9 Olefins
Olefins, also known as alkenes, are aliphatic hydrocarbons with one or more C–C
double bonds. They are very reactive compared to other hydrocarbon types, including
aromatics. They are commonly present in living organisms, but found only trace in a
few crude oils because they are usually reduced underground during burial to paraffins
with hydrogen or to thiols with hydrogen sulfide. Most of the olefins in refining are
generated by thermal processes. Olefins are important feedstocks for high polymers,
such as polyolefins and polyesters, also for gasolines through polymerization (poly
gasoline) and alkylation (alkylate). They are also important for making a wide variety
of chemicals. Figure 3.11 presents a few examples of olefins.
3.2.10 Alkynes
Alkynes,which contains triple bonds in the molecule, are not found in petroleum.
The simplest alkyne is acetylene which can be produced from partial combustion
of methane or natural gas with air or oxygen [7]. It is widely used as a fuel in
oxyacetylene welding and cutting metals. Most of alkynes are used as important raw
materials for making valuable chemicals and plastics.
The molecular composition of crude oil varies widely from formation to formation.
Four different types of hydrocarbon molecules appear in crude oil, shown in Table 3.5.
The relative percentage of each varies from oil to oil. The properties of the oil
are largely determined by the distributions of these compound types and individual
molecules.
Petroleum molecules can be lumped into groups, thereby allowing us to create
manageable models for physical properties, chemical composition and reaction kinet-
ics. Table 3.6 summarizes all of organic hydrocarbon molecules found in petroleum.
They are grouped into different compound classes and hydrocarbon types.
Crude oil also may contain a small amount of decay-resistant organic remains,
such as siliceous skeletal fragments, wood, spores, resins, coal, and various other
remnants of former life. Sodium chloride also presents in most crudes, and is usually
removed with desalting technology.
The crude oils in China are usually different than the conventional crude oils in other
countries, with the following unique properties [8]:
1. High viscosity, high wax content, high freezing point, relatively high specific
gravities between 0.85 and 0.95. By definition they are heavy oils.
2. Low atom ratio of H to C.
Table 3.5 Average and range of hydrocarbon types in crude oils [1]
Hydrocarbon Average (%) Range
Alkanes (paraffins) 30 15–60%
Naphthenes 49 30–60%
Aromatics 15 3–30%
Asphaltics 6 Remainder
3.4 Crude Oils in China 55
Table 3.7 Comparison of Chinese crude oils with crude oils outside China in sulfur and nitrogen
contents [8]
Crude oil in China Crude oil outside China
Location S, wt% N, wt% Location S, wt% N, wt%
Daqing 0.10 0.16 Canada (Athabasca) 4.8 0.4
Shengli 0.80 0.41 Texas (WTI) 0.34 0.08
Gudao 2.09 0.43 Saudi Arabia 2.55 0.09
Xingjiang 0.05 0.13 Venezuela (Boscan) 5.3 0.65
Table 3.8 Comparison of Chinese crude oils with crude oils outside China in metal contents [8]
Crude oil Metal content, µg/g Ni/V
Ni V Fe Cu
Daqing (China) 3.1 0.04 0.7 0.25 78
Shengli (China) 26.0 1.6 13.0 0.1 26
Gudao (China) 21.1 2.0 12.0 <0.2 11
Liaohe (China) 32.5 0.6 9.3 0.3 54
Huabei (China) 15.0 0.7 1.8 <0.3 21
Zhongyuan (China) 3.3 2.4 8.2 0.4 1.4
Iran 8.7 88.8 4.0 0.07 0.10
Saudi Arabia 11.1 31.4 1.9 0.06 0.35
Texas 5.8 20.8 5.8 0.4 0.28
UK 0.59 3.48 0.23 0.26 0.17
References
1. Hsu CS (1995) Hydrocarbons. In: Encyclopedia of analytical science, premiere edition, Aca-
demic Press, London, pp 2028–2034
2. Hsu CS (ed) (2003)Analytical advances for hydrocarbon research, Kluwer Academic/Plenum
Publishers, New York
3. Wikipedia: Petroleum. https://en.wikipedia.org/wiki/Petroleum. Retrieved 6 Aug 2015
4. Hsu CS, Lobodin V, Rodgers RP, McKenna AM, Marshall AG (2011) Compositional boundaries
for fossil hydrocarbons. Energy Fuels 25:2174–2178
5. Hsu CS, Dechert GJ, Robbins WK, Fukuda EK (2000) Naphthenic acids in crude oils charac-
terized by mass spectrometry. Energy Fuel 14(1):217–223
6. Barwise AJG (1990) Role of nickel and vanadium in petroleum classification. Energy Fuels
4:647–652
7. Watt LJ (1951) The production of acetylene from methane by partial oxidation.
Univ. of British Columbia. https://open.library.ubc.ca/cIRcle/collections/ubctheses/831/items/
1.0059187. Accessed 23 Nov 2018
8. Xu C, Yang C (2009) Petroleum refining engineering. Petroleum Industry Press, Beijing (in
Chinese)
Chapter 4
Classification and Characterization
Petroleum crude oils are broadly classified as paraffinic, asphaltic and mixed crude
oils. Paraffinic crude oils are composed of aliphatic hydrocarbons (paraffins), wax
(long-chain normal paraffin) and high-grade oils. Paraffinic crude oils can also con-
tain a small amount of asphaltic (bituminous) material. Traditional examples are
Pennsylvania grade crude oils. The paraffin wax in this crude oil is the origin of its
classification. Relative to other hydrocarbons, at a given temperature and pressure,
normal paraffins have a greater tendency to precipitate. This can cause problems
during oil production and processing. In the production process, waxy heavy paraf-
fins can build up in the tubing close the surface or they can block perforations.
This often happens in depleted reservoirs or reservoirs under gas-cycling conditions.
Isoparaffins are great for producing high-grade lubricating oils.
Asphaltic crude oils contain large proportions of asphaltic materials with low or
negligible concentrations of paraffins. It has been suggested that a crude oil should be
called asphaltic if the distillation residue contains less than 2% wax and paraffinic if it
contains more than 5% wax. Asphaltic crude oils are highly viscous, often appearing
in a sticky semi-solid or liquid state. They also have larger molecules and high boiling
points. Asphaltic oils are cheaper than paraffinic oils because they are heavier and
thicker. More effort is needed to convert them into fuels and other designated pur-
poses. Asphaltic crude is also known as pitch, bitumen, petroleum asphalt or asphalt.
Some asphaltic crude oils contain a predominance of cycloparaffins, which can be
feedstocks for high viscosity lubricating oils, which are more sensitive to tempera-
ture changes than paraffinic crudes. Such crude oils are also known as naphthenic
crude oils. Such crudes come from the U.S. Midcontinent region and parts of Iran.
Most crude oils are mixed, containing paraffins, naphthenes, and aromatic hydro-
carbons. These crudes usually have a lower API gravity and are heavier. Because of
their heaviness, naphthenic and aromatic crudes are worth less due to their higher
requirements for refinery processing. Naphthenic crudes may also contain metals
such as vanadium, arsenic, and iron. There are some crude oils which have up to
80% aromatic content, and these are used as aromatic-base oil.
Attempts have been made to define or classify petroleum based on various dis-
tillation properties when combined with another property such as density. In 1933,
Watson and Nelson [1] of Universal Oil Products (UOP, now a part of Honeywell)
introduced a ratio between the mean average boiling point and specific gravity that
could be used to indicate the chemical nature of hydrocarbon fractions and, therefore,
could be used as a correlative factor. UOP or Watson Characterization Factors (Kw)
are calculated with:
Kw = (TB )1/3 /G
Fig. 4.1 Typical characterization factors versus various crude oil API gravities [2]
4.1 Petroleum Classification 59
where TB is mean average boiling point in °K (°C + 273) and G is specific gravity
at 60 °F (15.6 °C).
CI is an indicator of the aromaticity of a crude oil. The CI scale ranges from
zero for straight-chain paraffins to 100 for benzene. The CI of a pure compound
can be calculated from its boiling point and specific gravity, shown in Fig. 4.2 for
a few representative compounds. In the figure, the correlation lines are drawn along
with reciprocal of boiling point as y-axis and specific gravity at 60 °F as x-axis.
Normal paraffins are along with the CI = 0 line. Aromatic compounds with specific
gravity similar to benzene are distributed between CI = 20 and 100 where benzene
is located. Naphthenes are in between paraffins and aromatics, with CI between 10
and 50. Cyclohexane has a CI of 51, while polycyclic aromatics have CI higher than
100.
For crude oils, CI values between 0 and 15 indicate a predominance of paraffinic
hydrocarbons; values from 15 to 50 indicate a predominance either of naphthenes
or of a mixture of paraffins, naphthenes, and aromatics; values above 50 indicate a
predominance of aromatics.
60 4 Classification and Characterization
Both the Watson Characterization Factor and Correlation Index are solely based
on two physical properties: average boiling point and specific gravity (density). They
do not define the relative proportions of open chain paraffins and ring naphthenes
(both are saturates) and aromatics, the actual compound types. There certainly are
overlaps between these compound types, thus, in crude oils of different categories
determined by these two methods. For example, the high specific gravity may be
due to both naphthenic and aromatic compounds as well as asphaltic and resinous
materials. Their chemical nature may be quite different.
A chemical classification was proposed by Tissot and Welte based on the oil con-
tents of various hydrocarbon types in a ternary plot [5]. The types include alkanes
(normal and iso), cyclo-alkanes (naphthenes) and aromatics including heteroatom-
containing aromatics (NSO compounds), resins and asphaltenes, as shown in Fig. 4.3.
The oils are defined into six classes: paraffinic oils, naphthenic oils, paraffinic-
naphthenic oils, aromatic-intermediate oils, aromatic- naphthenic oils and aromatic-
asphaltic oils. They truly reflect the type of hydrocarbons that oil contains. Paraf-
finic oils have open-chain hydrocarbon content of greater than 50%. In naph-
thenic oils, saturated cyclic compounds account for >50% naphthenes. Paraffinic-
naphthenic oils contain more than 50% saturated molecules and aromatics account
for 25–40% with resins and asphaltenes vary between 5 and 15%. In aromatic-
intermediate oils, aromatics are 40–70%, resins and asphaltenes constitute 10–30%
of the oil. Aromatic-naphthenic oils are biodegraded oils originated from paraf-
finic and paraffinic-naphthenic oils. Aromatic-asphaltic oils are mostly biodegraded
aromatic-intermediate oils, with a few true aromatic oils. They are heavy and vis-
cous, with resins and asphaltenes content varies between 30 and 60%. The variation
trends of hydrocarbon contents due to thermal maturation and biodegradation of the
oils are also indicated by the arrows.
There are other ways to classify the crude oil, including:
(1) Classification according to API gravity:
Light—API gravity >32
Medium—API gravity 20–32
Heavy—API gravity 10–20
Extraheavy—API gravity ≤10, viscosity <10,000 cP
Bitumen—API gravity ≤10, viscosity >10,000 cP.
Figure 4.4 shows the classification of heavy oils with API gravity <20. The API
gravity for both Extra heavy oils and Bitumen is less than 10. Bitumen’s viscosity is
greater than 10,000 cP, while Extraheavy oil’s viscosity is less than 10,000.
(2) Classification according to sulfur content:
Sweet—<0.5% S,
Intermediate—0.5–1.5% S
Sour—>1.5% S.
Typical sweet crudes include West Texas Intermediate (WTI, which is a so-
called benchmark crude traded on the New York Mercantile Exchange) from mostly
4.1 Petroleum Classification 61
Fig. 4.3 Crude oil chemical classification based on hydrocarbon types [5]
Louisiana and Oklahoma, Nigerian crudes, and Brent crude oils from the North Sea
(Brent is a benchmark crude traded on the International Petroleum Exchange). Sour
crudes include Alaska North Slope, Venezuelan, and West Texas Sour from fields
like Yates and Wasson. Intermediate crudes include California Heavy, such as from
the San Joaquin Valley, and many Middle East crudes.
The complex mixtures which comprise crude oil are separated by distillation and sol-
ubility. Solubility depends largely on polarizability and polarity. A conventional pro-
cedure is to separate petroleum into saturates, aromatics, resins and asphaltenes frac-
tions, known as a SARA (Saturates-Aromatics-Resin-Asphalt) analysis. Asphaltenes
could be pentane, hexane or heptane insoluble. Lighter solvent is preferred for lighter
oils to remove asphaltenes. Saturates and aromatics are determined by adsorption
(column) chromatography on alumina or silica gel. Resins are recovered by polar
solvents after the collection of Saturates and Aromatics. There are several analytical
methods for this, including gravity-driven chromatographic separation, thin layer
chromatography (TLC), and high-performance liquid chromatography (HPLC). The
results from these analytical techniques do not produce identical results. Hence, they
are not interchangeable. For comparison purposes, the same techniques should be
applied in the same order every time. The fastest and least expensive method is
Iatroscan TIC with a flame ionization detector (FID) to determines the weight per-
cent of each fraction in the oil. The sample is spotted onto a quartz rod coated with
sintered silica particles and separated over thin layer surface. The rod is then scanned
directly in the FID at rated speed for quantification.
In theory, SARA analysis is applicable to all types of crudes, However, most of the
problems in SARA analysis arise from the loss of light ends. Hence, SARA analysis
should be confined to heavy oils, or heavier fractions of the oils with lighter fractions
determined by other methods, such as gas chromatography (GC). Figure 4.5 shows
SARA analytical scheme for crude oil with the fraction boils above 210 °C. The light
fraction that boils below 210 °C is distilled off for subsequent GC analysis that can
be used to determine the saturates and aromatics contents. Resin and asphaltene are
not present or just in negligible amounts in light ends. The fraction that boils above
210 °C is treated with n-hexane in the example shown to precipitate off asphaltene.
However, the solvent can also be n-pentane for light crudes or n-heptane for heavy
crudes, which may be a better choice because n-hexane may contain azeotropic
impurities, isohexanes and cyclohexane that are difficult to remove. The n-hexane
insoluble precipitate is Asphaltene, which is subjected to infrared or nuclear magnetic
resonance spectroscopy for structure determination. The n-hexane soluble fraction,
i.e., deasphaltened oil (DAO) or maltene, is subjected to liquid chromatography sepa-
ration on alumina or silica gel into saturates, aromatics and resins fractions. Saturates
elute off the chromatography column with n-hexane. Benzene or toluene is normally
used to elute off aromatics. A polar solvent, such as methanol or pyridine or a mixed
4.2 SARA Analysis for Characterization 63
Crude oil
n-heptane precipitation
1:1 methanol:
n-hexane toluene
dichloromethane
Table 4.1 Typical SARA distributions of light, medium and heavy crude oils
Class of crude oil Saturates Aromatics Resins Asphaltenes
Light Crude 92 8 1 0
Medium Crude 78 15 6 1
Heavy Crude 38 29 20 13
content can change during production or over time; hence, remeasurements would
become necessary.
The saturates fraction consists of nonpolar compounds including linear (normal
paraffins), branched (isoparaffins) and cyclic (cycloparaffins) saturated hydrocar-
bons. Aromatics are more polarizable, containing one or more aromatic rings. Both
resins and asphaltenes have polar constituents. Resins are miscible with light n-alkane
(n-C5 , n-C6 or n-C7 ), while asphaltenes are insoluble in even an excess of n-alkane
solvent. Asphaltene molecules are colloidally dispersed in crude oil are stabilized by
polar molecules of aromatic and resin molecules, preventing any major aggregation
of the asphaltene [8].
Asphaltene precipitation is caused by a number of factors including changes in
pressure, temperature and liquid-phase composition. Hence, the composition and
structure of asphaltene in reservoir and laboratory conditions would be different.
The equilibrium between the reservoir crude oil and asphaltene as self-assembled
nanocolloidal particles would be destroyed by releasing dissolved gas upon drilling.
Drilling, completion, stimulation, and hydraulic fracturing activities can also induce
precipitation in the near-wellbore region.
References
1. Watson KM, Nelson EF (1933) Improved methods for approximating critical and thermal prop-
erties of petroleum. Ind Eng Chem 25(8):880–887
2. PetroWiki: Crude characterization. http://petrowiki.org/Crude_oil_characterization. Retrieved
9 Oct 2017
3. Smith HM (1940) Correlation index in crude oil analysis, U. S. Bureau of Mines, Tech. Paper
610
4. (a) Smith HM (1960) Correlation index in crude oil analysis, U. S. Bureau of Mines, Tech. Paper
610, 1940; (b) Gruse WC.; Stevens DR. Chemical Technology of Petroleum, McGraw-Hill
5. Tissot BP, Welte DH (1984) Petroleum formation and occurrence, 2nd edn. Springer-Verlag,
Berlin-Heidelberg-New York-Tokyo
6. Qian K, Hsu CS (1992) Molecular transformation in hydrotreating processes studied by online
liquid chromatography mass spectrometry. Anal Chem 64:2327–2333
7. Hsu CS, Hendrickson CL, Rodgers RP, McKenna AM, Marshall AG (2011) Petroleomics:
advanced molecular probe for petroleum heavy ends, special feature: perspective. J Mass Spec-
trom 46:337–343
8. Aske N, Kallevik H, Johnsen EE, Sjoblom J (2002) Asphaltene aggregation from crude oils and
model systems studied by high-pressure NIR spectroscopy. Energy Fuels 16:1287–1295
Part II
Exploration and Production
of Petroleum—Upstream
Science and Technology
Chapter 5
Petroleum System and Occurrence
There are several terms one must first understand when studying a petroleum system.
The first term is basin. A basin (sedimentary basin or petroleum province) is an area
of the Earth where there is a large depression. This depression allows for the space to
be filled with sediments and oils. Next is stratum. When there is layering of rocks in
the ground this is referred to as a stratum. A Reservoir is an underground collection
of oil or gas within porous rock formations. A play is a region that when studied
shows a specific source, reservoir or trap that has potential for hydrocarbon build up.
A prospect is an individual exploration target where a specific trap has been
identified and mapped but has not been drilled yet. A reserve is an accumulation in
which the presence of oil and gas has been verified by drilling. An outcrop is bedrock
that is visible above ground to show ancient superficial deposits. Being able to see
the bedrock on the surface of the Earth simplifies determination of the porosity and
CO2
Total Organic
Carbon in Living
Organizations
1012 tons
Petroleum
1011 tons
permeability of the rock. These two properties are the most important to the discovery
of oil. Porous rocks such as sandstone, limestone, or dolomite allow petroleum to
build up in the reservoir, but also allow oil and gas to migrate. Impermeable rocks
such as clay or shale trap oil and gas, inhibiting large-scale accumulation and also
preventing further migration.
Figure 5.2 shows the relationship between the geological scales for petroleum
exploration. Once the prospects are identified, the next step is to lease the target
areas for drilling wells for production.
5.2 Petroleum System 69
Fig. 5.2 Relative geological scales for petroleum exploration and production [10]
Crude oil and natural gas come from organic-rich sediments, which are formed
by the accumulation and subsequent burial of dead organisms deep under the earth’s
surface in stratified layers. The stratified sediment bed is also referred to as the
formation. The rock above the organic-rich source rock is called the overburden,
which is sedimentary rock that encases the source rock, seal rock, and reservoir
to provide thermal stress. Under thermal stress by overburden pressures of several
hundreds or thousands of atmospheres over several hundred million years, the organic
material is converted and compacted into kerogen. Oil and natural gases are released
from the kerogen then migrate through porous carrier rock, such as sandstone, or
along fault lines, to accumulate under impermeable cap rock or salt domes which
functions as seal or trap (reservoir). The raw crude oil in the reservoirs can be mixtures
of liquid and gas trapped in rock and sand and other materials.
The petroleum system, as represented in the simplified picture shown in Fig. 5.3,
encompasses porous source rock containing kerogen, which is the ultimate source of
petroleum and natural gas, permeable carrier rock for oil and gas migration from the
source to the reservoir, porous reservoir rock and impermeable seal rock (also known
as cap rock) for accumulations, and overburden. Source rock and reservoir rock are
also permeable. A common transport mechanism is simple diffusion driven by buoy-
ancy and ground-/pore-water flow. All these generation-migration-accumulation pro-
cesses must be present at the right time and right place for petroleum formation and
accumulation.
70 5 Petroleum System and Occurrence
overburden
anticlinal seal
trap reservoir oil migration
source gas migration
underburden
+ + basement stratigraphic
trap
Fault-bound
traps
Since the Renaissance, scientists have debated two theories on the formation of coal,
petroleum and other fossil hydrocarbons: inorganic through abiogenic processes deep
within the Earth versus organic from sediments through paleotransformation during
burial [2]. The inorganic-origin hypothesis implied that petroleum appeared on Earth
before life.
Scientists had put forward inorganic theories based on the following evidence:
1. Some inorganic reactions generate a certain amount of hydrocarbons.
2. The gas and lava which discharge from volcanos contain hydrocarbons.
3. Many planets contain hydrocarbons. For example, the atmospheres of Jupiter
and Uranus contain methane.
In 1953, the historic Miller-Urey experiment showed how simple molecules can
be transformed into amino acids, the building blocks of proteins. Miller introduced
water, methane, ammonia, and hydrogen into a sterile 5-liter glass flask connected
to a smaller flask half filled with liquid water [3]. The water in the smaller flask was
heated to generate steam, which was allowed to enter the larger flask. Continuous
electrical sparks were fired between two electrodes in the larger flask to simulate
lightning. The larger flask was cooled so that the water condensed and trickled into
a U-shaped trap at the bottom of the apparatus. After a day, the solution in the trap
turned pink. After one week of continuous operation, the flask was cooled, and the
liquid in the trap was quickly dosed with mercuric chloride to prevent microbial
contamination. Paper chromatography revealed the presence of glycine, α- and β-
alanine.
Miller’s original experiments yielded 20 different amino acids. Using variations of
the Miller experiment, in addition to amino acids, the other major building blocks of
life, sugars, nucleic acid bases, and lipids, also have been produced [4]. It is believed
that from these building blocks, primitive life evolved.
5.3 Origin and Occurrence 71
Fig. 5.4 Comparison of Chlorophyll a in plants (a) with porphyrins (b and c) found in petroleum
[6]
In the 18th century, fossil evidence led scientists to conclude that coals were
derived from plant remains. These conclusions influenced scientists to look into
similar explanations for the origin of petroleum. Mikhail Lomonosov gets credit for
suggesting that petroleum and bitumen were produced underground from coal at
high pressure and temperature [5].
Modern theories, in which petroleum was said to originate from ancient sedimen-
tary, organic-rich rocks, emerged in the 19th century [2]. In 1863, T. S. Hunt of the
Canadian Geological Survey concluded that bitumen in some North American Pale-
ozoic rocks resulted from paleo-transformation of ancient marine microorganisms.
Further scientific advances in paleontology, geology, and chemistry led to refinement
of the model. In 1936, Alfred Treibs linked the chlorophyll in plants to the porphyrins
in petroleum, shown in Fig. 5.4 [6]. Additional geochemical evidence was provided
by discovery of optical activity in moderately mature oils [7]. Further, there were
a host of hydrocarbons that could be traced back to specific biological precursors,
known as fossil fuel biomarkers [8].
72 5 Petroleum System and Occurrence
Based on the evidence from radiometric age dating, physicists have deduced that
the age of the Earth is 4.54 ± 0.05 billion years (4.54 × 109 years ±1%) [9]. This
calculation encompasses measuring the concentration of the stable end product of a
radioactive decay chain, coupled with knowledge of the half-life and initial concen-
tration of the decaying isotope. In this instance, the end product is lead (specifically
206
Pb) and the decaying isotope is uranium (238 U). The latter has a half-life of 4.47
billion years and persists in stable minerals, making it ideal for dating the Earth.
U/Pb dating is usually performed on the mineral zircon (ZrSiO4 ), which incorpo-
rates uranium and thorium atoms in its crystalline structure but strongly rejects lead.
Therefore, one can assume that the entire lead content of the zircon is radiogenic.
Similar measurements are made on meteorites,
Figure 5.5 depicts the major division of geological time. The Eon is the largest
division of the geologic time scale, covering time intervals of about 500 million years
to more than a billion years [10]. The 4 eons in Earth history are:
Hadean, ~4.6 billion to 4 billion years ago;
Archean, 4 billion to 2.5 billion years ago;
Proterozoic, 2.5 billion to 545 million years ago; and
Phanerozoic, 545 million years ago to present.
When Earth was newly formed, the surface was molten and constantly bombarded
by meteorites in the Hadean eon. The moon was formed during this eon. As the
planet cooled in the Archeon eon, an atmosphere formed, largely from gases spewed
from volcanoes. The gases included water vapor, hydrogen sulfide (low temperature)
croscopic size and lack of hard part, bacteria are rarely fossilized. Their geological
records are often associated with plant tissues and animal and insect remains.
The autotrophic phytoplankton and heterotrophic zooplankton are foundation of
the food chain within the pyramid of life. High abundances of fossil phytoplank-
ton groups, including blue green algae, green algae and arcritarchs, appeared in the
Paleozoic era, particularly in Ordovician, Sulurian and Devonian periods [11]. More
abundant phytoplankton groups, also including dinoflagellates, coccolithophorids,
silicoflagellates, euglenids, diatoms, ebridians and discoasters, appeared in the Meso-
zoic and Cenozoic eras [11].
5.5 Paleo-Transformation
More evidence has been obtained to increase our understanding of how biological
molecules were transformed into petroleum molecules. Upon death, some micro-
organisms, such algae and plankton, which are abundant in ocean, can escape full
oxidation (create a leak in carbon cycle) with quick settlement and deposition of
debris in layers. These sediments can be pushed more deeply under the surface by
tectonic movement and undergo molecular transformation due to thermal stress under
anoxic conditions. Most oil found under dry land today was below sea level when
the organic deposition occurred. Subsequent tectonic movement lifted and folded the
rock into its present morphology [12].
When dead organic matter (diatoms, planktons, pollens and spores) is buried
below about 200 m, the organisms decompose in the presence of water and miner-
als in a process called diagenesis. This includes chemical, physical and biological
changes to the sediment. The organic matter breaks down into proteins, lipids and
carbohydrates (biopolymers) due to increasing temperature and pressure. These are
further broken down into amino acids that in turn become fulvic, humic, and humin
acids (geopolymers, precursors of kerogen). This transformation in the first few hun-
dred meters of burial occurs at temperatures around 60 °C. Eventually, the decaying
5.5 Paleo-Transformation 75
The ubiquitous presence of biological markers (biomarkers) in crude oils and coals
provides the evidence of the biological origin of petroleum and coal. Biomarkers are
the molecules that retain the basic carbon skeleton of biological precursors. Crude oil
compositions vary widely depending on the oil sources, the thermal regime during oil
generation, the geological migration and the reservoir conditions. Most of the crude
76 5 Petroleum System and Occurrence
oil constituents undergo changes in their chemical structure over time as an effect
of several factors, among which are biodegradation and weathering. Biomarkers are
more degradation-resistant in the environment than other hydrocarbon groups in oil.
Biomarkers are very useful molecules for exploration, exploitation and produc-
tion in upstream applications. Some representative biomarkers and their biological
precursors are shown in Fig. 5.8. With detailed studies, biomarkers can be used for
assessment of source input, age, maturity and alteration of petroleum [13, 14]. They
can also be used as fingerprinting for oil-oil and oil-source rock correlation during
exploration to determine the source of oil, source potential and migration pathways
of a source rock to the reservoir. In environmental applications, biomarker analy-
sis is of great importance in forensic investigations. It is used for determination of
the source of spilled oil, differentiation and correlation of oils, and monitoring the
degradation process and weathering state of oils under a wide variety of conditions
[15].
5.7 Kerogen
Upon death and burial, living organisms, such as planktons, algae, spores and pol-
lens, begin to undergo decomposition or degradation without full oxidation. Large
biomolecules, such as proteins and carbohydrates, begin to break down. The dis-
mantled components can then condense, accompanied by the formation of mineral
components, to form geopolymers. The smallest units are fulvic acids, the medium
units are humic, and the largest units are humins. The subsequent sedimentation and
5.7 Kerogen 77
boxylic acid functional groups. The material is thick, resembling wood and coal. It
tends to produce coal and gas.
Type IV (residue) kerogen contains mostly decomposed organic matter in the
form of highly condensed polycyclic aromatic hydrocarbons (PAH’s). The H:C ratio
is <0.5 and it has no potential to produce hydrocarbons.
Carbon has three isotopes: 98.89% 12 C, 1.11 wt% 13 C and trace (~10−12 ) 14 C in
natural abundance. 14 C is radioactive while the former two are stable. 14 C is pro-
duced by reaction of thermal neutrons from cosmic radiation with 14 N in the upper
atmosphere. It enters biosphere as CO2 via photosynthesis. By emitting a β particle
(an electron), 14 C radioactively decays to form stable 14 N.
14
C(6p + 8n) − β →14 N (7p + 7n)
Since dead tissue does not absorb 14 CO2 , 14 C can be used for radiometric dating
of dead biological materials, such as Egyptian mummies. However, the half-life of
14
C is 5570 years, so it essentially disappears after 5 half-lives, or ~30,000 years.
The age of natural gas and crude oils is on the order of several hundred million years;
hence, 14 C cannot be used for the age determination of natural gas and crude oils.
The age of petroleum is based on radioactive materials with much long half-lives,
such as uranium, in fossils or rocks in close proximity.
As mentioned earlier, biogenic methane from recent decay of biological materials
contains radioactive 14 C that is absent in the methane in natural gas and reservoirs.
For most stable isotopes, the magnitude of fractionation from kinetic and equilib-
rium fractionation is very small; for this reason, enrichments are typically reported in
“per mil” (‰, parts per thousand). The ratio of the other two stable carbon isotopes,
13 12
C/ C, is widely used in geochemistry to distinguish between organic and inorganic
carbon and to determine sources of gas and oil. It is expressed as the enrichment of
13
C (δ13 C), which is the 13 C/12 C ratio compared to that of a standard, shown below:
δ13 C(per mil, 0/00) = 13 C/12 C sample / 13 C/12 C standard − 1 × 1000
The commonly used standard reference material is belemnite from the PeeDee
Formation in South Carolina (PDB with 13 C/12 C = 1.123 × 10−2 ) or Solenhofen
limestone (NBS-20 with 13 C/12 C = 1.1218 × 10−2 ). The Selenhofen limestone is is
more commonly used now due to the depletion of PDB.
The 13 C/12 C ratio can be obtained from isotope ratio mass spectrometry, shown
in Fig. 5.10, with a precision of 1 part in 10,000. In the analysis, the sample is
combusted to CO2 , which is ionized by an electron beam generated by filament at
70 eV energy; results are highly repeatable. The resulting ions are repelled from the
80 5 Petroleum System and Occurrence
Ion Source
ion source and accelerated into a magnet. The ions of different mass-to-charge (m/q)
ratios are deflected at different angles. Three Faraday cage detectors are placed to
measure the abundances of 12 CO2 (m/q = 44), 13 CO2 (m/q = 45) and 14 CO2 (m/q =
46). This gives the 13 C/12 C ratio, which is converted into δ13 C and compared to the
standard.
The δ13 C value for the bicarbonate in seawater is about the same as that for PDB.
The CO2 in air has 7% less 13 C than PDB. The U.S. National Bureau of Standards
(NBS) has an oil standard with a δ13 C of −19.4‰ on the PBD scale, which can be
used as a secondary standard. The carbonates rocks have a δ13 C range between +
5‰ and −5‰. The δ13 C ranges of marine plankton and terrestrial plants are in the
ranges of −14 to −28‰ and −14 to 32‰, respectively. Fossil organic matter falls in
the range covered by marine plankton and terrestrial plants. Petroleum is typically
between −18 and −34‰, another indication of organic derived material.
The δ13 C the biomethane ranges from −55 to −85‰, while for methane formed
by thermal cracking at depth (geological or fossil methane) the range is from −25
to −60‰.
All crude oils and natural gases contain sulfur compounds. Sulfur is essential for life.
It is a constituent of many proteins, enzymes and cofactors. The incorporation of sul-
fur into fossil hydrocarbons can come from both organic and inorganic sources [18].
From dead living organisms, mineralization converts organic sulfur into inorganic
forms, such as H2 S, elemental sulfur, polysulfides and sulfide minerals. Oxidation
5.9 Sulfur Incorporation into Fossil Hydrocarbons 81
References
1. Ourisson G, Albrecht P, Rohmer M (1984) The microbial origin of fossil fuels. Sci Am
251:44–51
2. Walters CC (2006) The origin of petroleum. In: Hsu CS, Robinson PR (eds) Practical advances
in petroleum processing. Springer, New York
3. Miller SL (1953) Production of amino acids under possible primitive earth conditions. Science
117(3046):528–529
4. Horowitz H (1992) Beginnings of cellular life. Yale University Press, New Haven
5. http://pineislandnews.com/Pine_Island_News_Blog/content/1757-lomonosov-fathered-
theory-oil-originates-biological-material. Retrieved on 10 Nov 2016
6. Treibs A (1936) Chlorophyll and hemin derivatives in organic mineral substances. Angew
Chem 49:682–686
7. Oakwood TS, Shriver DS, Fall HH, Mcaleer WJ, Wunz PR (1952) Optical activity of petroleum.
Ind Eng Chem 44:2568–2570
8. Philp RP. C&E News, 2/10/86, pp 28–43
9. Patterson C (1956) Age of meteorites and the earth. Geochim Cosmochim Acta 10(4):230–237
10. Ali HN (2017) Fundamentals of petroleum geology for exploration. In: Hsu CS, Robinson PR
(eds) Springer handbook of petroleum technology. Springer, New York
11. Tissot BP, Welte DH (1984) Petroleum formation and occurrence, 2nd edn. Springer-Verlag,
Berlin-Heidelberg-New York-Tokyo
12. Hunt JM (1979) Petroleum geochemistry and geology. W. H. Freeman and Company, San
Francisco
13. Hsu CS, Walters CC, Isaksen GH, Schaps ME, Peters KE (2003) Biomarker analysis in
petroleum exploration. In: Hsu CS (ed) Analytical advances for hydrocarbon research Kluwer
Academic/Plenum Publishers, New York/Boston/ Dordrecht/London/Moscow
14. Peters KE, Moldowan JM (1993) The biomarker guide: interpreting molecular fossils in
petroleum and ancient sediment, Prentice Hall
15. Prince RC, Elmendorf DL, Lute JR, Hsu CS, Haith CE, Senius JD, Dechert GJ, Douglas GS,
Butler EL (1994) 17*(H),21*(H)-Hopane as a conserved internal marker for estimating the
biodegradation of crude oil. Envrion Sci Tech 28:142–145
82 5 Petroleum System and Occurrence
16. Peters, K. E.; Schenk, O.; Scheirer, A. H.; Wygrala, B.; Hantschel, T (2017) Basin and Petroleum
System Modeling. In: Hsu CS, Robinson PR (eds) Springer handbook of petroleum technology,
Springer, New York
17. Walters CC (2017) Origin of petroleum. In: Hsu CS, Robinson PR (eds) Springer handbook of
petroleum technology. Springer, New York
18. Ivanov MV and Freney JR (1983) The global biogeochemical sulphur cycle. Chichester, John
Wiley
19. Wikipedia: Isotope-ratio mass spectrometry. https://en.wikipedia.org/wiki/Isotope-ratio_
mass_spectrometry. Retrieved on 10 Nov 2016
Chapter 6
Exploration for Discovery
In modern times, a required prerequisite to looking for oil is obtaining the rights to
do so. Leases must be acquired from land- or mineral-rights owners, and licenses
must be obtained, often from multiple legal jurisdictions—local, regional, state or
provincial, and/or national governments. Environmental, health and safety standards
must be understood and met. To develop large offshore basins can cost tens of bil-
lions of dollars, more than even the biggest oil companies can afford on their own.
Several platforms might be required, and the cost of a platform in a severe-weather
locale, such as the North Sea, can exceed $2 billion; in 1976, the ill-fated Piper
Alpha platform cost $1.6 billion—equivalent to about $3 billion in 2015. Therefore,
partnerships are formed to share capital costs, risks, and eventual profits.
There are generally two distinct phases in searching for natural resources:
prospecting and exploration. Prospecting is the search for unknown deposits. Explo-
ration follows this up with precise investigations and development of the reserves
and deposits found. Production can only begin after exploration has demonstrated
that sufficient amounts of resources can be extracted.
In ancient times, oils and gas were discovered near natural seeps or leaks to
the Earth’s surface. Such easy discoveries do not occur in modern days. Various
geophysical methods have to be applied in the search for oil and gas reservoirs of
commercial value. As discussed in the previous chapter, definitions applicable to
petroleum exploration include the following:
• A Basin is a large-scale depression on Earth’s surface filled with sediments and
oils. Basins can change over geological time due to movement of the Earth driven
by plate tectonics. The majority of oils in Middle East are marine-sourced. The
desert in Saudi Arabia today was under the ocean several hundred million years
ago. Similarly, marine fossils are found high in Himalayas. On the other hand,
many of today’s offshore oils are terrigenous-sourced.
• A Play is a group of oil fields or prospects in the region that are controlled by the
same set of geological circumstances with conditions favorable for hydrocarbon
accumulation.
• A Prospect is an individual exploration target which can be a specific trap that has
been identified and mapped but has not been drilled yet.
In the petroleum system described earlier, outcrops provide visible evidence of
alternating layers of porous and impermeable rocks. Porous rock (typically sandstone,
limestone, or dolomite) provides the reservoir and migration media of petroleum.
Impermeable rock (typically clay or shale) acts as a trap to prevent migration.
A book by Charles F Conaway presents a good overview of how we look for
and produce petroleum [1]. Commercially significant petroleum systems include the
following:
• Organic-rich source rock in which oil and gas were formed.
• Pathways comprised of permeable rock such as sandstone and limestone, or cracks
in faults, which allow oil and gas to migrate. Often, the movement of oil and gas
is promoted by the flow of underground water.
• Porous, permeable reservoir rock capped by impermeable rock, which prevents
fluids from migrating to the surface. Shale and salt domes are common caps.
When petroleum geologists look for oil and gas, they search for geological struc-
tures that might serve as traps. In so doing, they rely heavily on reflection seismology.
During seismic exploration, explosives or thumper devices send sound waves through
the earth. Reflected sound waves are measured with hydrophones (in water) and
geophones (on land). Different layers reflect sound in different ways. With the help
of sophisticated software, geophysicists transform seismic data into 3-dimensional
maps that show the structure of subsurface rock formations.
For petroleum accumulation, there are four major kinds of reservoir traps, three
of which (structural traps) are illustrated in Fig. 6.1.
• Anticline traps are, in essence, inverted bowls. They can be symmetrical or asym-
metrical. A steep anticline might be called a dome. Anticline traps hold most of
the world’s conventional crude oil.
• Faults are formed at the boundaries of cracks in the earth’s crust. The four major
kinds of faults are thrust, lateral, normal and reverse. Thrust and lateral faults
are created by horizontal movement. The San Andreas Fault in California is a
well-known example of a lateral fault. To be more specific, the San Andreas is
a right-lateral, strike-slip fault. Normal and reverse faults are created by vertical
movement and are more likely to create fault traps for petroleum. The fault acts
as a trap when it pushes an impermeable substance (such as clay or shale) across
the fault line through permeable rock, blocking fluid migration.
• A salt-dome trap is created when a mass of underground salt in an evaporite from
a former ocean is pushed up by buoyancy and underground pressure into a dome.
The salt dome breaks through layers of rock and pushes them aside as it rises. The
salt is impermeable, and if it abuts porous rock, it can serve as a reservoir cap.
6.1 Exploration (Looking for Oil and Gas) 85
Fig. 6.1 Three types of hydrocarbon traps—anticline (a), fault (b), and salt-dome (c). As with oil
and gas, the water layers are not pools, but water-saturated porous rock
86 6 Exploration for Discovery
Porosity and permeability are the two most important properties of reservoir rocks.
Porosity is the fraction of volume that is not occupied by the solid framework of the
rock. It may be classified either as absolute or effective. Absolute porosity is the total
void volume relative to the bulk volume regardless of whether or not the void volumes
are interconnected. Without interconnections, no fluid conductivity (permeability)
would occur, leading to low effective porosity. Effective porosity is the ratio of
connected pore volumes to the bulk volume, including interconnections. However,
not all the interconnected pores may be available for the storage or transmission of
oil. Some pores may be filled with water which cannot be replaced by oil, and some
oil-filled pores may not allow the oil to move out. Only solution gas drive can force
such oil out of these pores.
In oil reservoirs, the porosity represents the percentage of total void space which
can be occupied by oil or gas. Hence, it determines the maximum storage capacity
of the rock for oil or gas. The porosities of the most frequently found reservoirs are
between 10 and 20%.
Permeability is the ability of a rock to allow a fluid to flow through its intercon-
nected pores. Both porosity and permeability are affected by grain size, grain dis-
tributions, angularity, the amount of cementation, and consolidation. The presence
of porosity does not ensure the rock will have a significant permeability, because
permeability depends on the shape, size and degree of interconnection of the pores.
However, for a rock to have permeability, it must be porous.
Outcrops which expose bedrock or ancient superficial deposits are visible on the
surface of the Earth. They provide visual evidence of alternating layers of porous
and impermeable rocks. Porous rock (typically sandstone, limestone, or dolomite)
provides the reservoir of petroleum and migration routes for oil and gas while imper-
meable rock (typically clay or shale) acts as a trap to prevent oil and gas from
migration for accumulation.
For oil shale, reservoir rock and source rock are the same because the oil has not
migrated. Permeability is practically zero and the exploitation is only possible by
mining or hydraulic fracturing.
6.3 Geophysical Methods 87
Initial attempts to locate petroleum after its initial discovery were limited to above-
ground efforts. Areas where petroleum had begun to reach the surface were easy
first targets for drilling. In the mid 1800s, it was discovered that certain formations
above the ground, such as broad low hills, corresponded to oil reserves. Such features
were formed where the Earth had folded, so there was a greater chance of having
oil beneath them for the reasons explained in the section on the formation of traps.
Still, drilling didn’t always find oil, and success rates were very low. Nowadays,
above ground techniques are much more sophisticated. Instead of looking at the rock
formations from the ground level, aircraft and satellites are used to survey the earth.
Aircraft can also measure slight changes in the gravitational force above the earth,
allowing geologists to guess at the density of what lies below the surface.
Because drilling to the depths where oil or gas is found is very costly, companies
must perform several tests in order to justify drilling in certain prospects. Geophysical
methods applied for oil and gas discovery include gravity, seismic, magnetic, electric,
electromagnetic and borehole logging (or well logging).
In large scale remote areas, gravity survey methods are often used to determine
where a seismic survey should be acquired. The gravity method is a non-invasive,
non-destructive remote sensing geophysical technique that measures densities of
rocks underground at specific locations for a better understanding of the subsurface
geology. The higher the gravity value, the denser is the rock beneath. No energy
needs to be put into the ground in order to acquire data.
Seismic exploration is the search for commercially economic subsurface deposits
of crude oil, natural gas and minerals by the recording, processing, and interpretation
of artificially induced shock waves in the earth. Seismic surveys use the propagation
of natural or artificial seismic waves in the subsurface to measure the properties of
rocks and fluids and to locate likely rock structures underground in which oil and gas
might be found [2]. Seismic waves are generally thought of as having low frequency
(in the range of five to eight Hertz) and large amplitude.
Artificial seismic energy is generated on land by vibratory mechanisms mounted
on specialized trucks to fire shock waves into the ground. Seismic waves reflect and
refract off subsurface rock formations and travel back to acoustic receivers called
geophones. It uses the basic relationship between the velocity and travel time (mea-
sured in milliseconds) of sound waves to estimate distance (depth) of subsurface
structures. A variety of other wave properties is used to characterize rocks and fluids.
The travel times of the returned seismic energy, integrated with existing borehole well
information, aid geoscientists in estimating the structure (folding and faulting) and
stratigraphy (rock type, depositional environment, and fluid content) of subsurface
formations, and determine the location of prospective drilling targets by identifying
underground structures conducive to oil accumulation.
Although seismic lines are two-dimensional, in other words they represent a cross-
section, it is possible these days to collect 3D seismic and 4D seismic data as well,
where 3D data represents a volume, and 4D data includes the dimension of time. For
88 6 Exploration for Discovery
Geological Structure
Fig. 6.2 Seismic data acquisition (a) and resulting 2D-map for interpretation (b)
• Nuclear Magnetic Resonance (NMR) Logs—The logs use the Earth’s magnetic
field as the external field and electromagnets for the locally imposed field. It is
used for the determination of the presence and quantities of different fluids (water,
oil and gas), pore size distribution, effective porosity and permeability, bound and
free water saturation, etc. However, NMR logs are less sensitive to matrix lithology
than conventional (acoustic, density and radioactivity) logs.
Petroleum geochemistry deals with the application of chemical principles to the study
of the origin, generation, migration, accumulation, and alteration of petroleum [5].
Fig. 6.4 depicts the basic concept behind geochemistry for oil exploration [6]. If oil
from Well 1 correlates with organic extracts from shale in Well 2, the fault is not
acting as a seal and oil migrates through the fault. If no correlation is found, the fault
may be acting as a trap for oil formed from shale in Well 2, and therefore, it is a
target for further drilling. If no oil is found in Well 2, the extract of the source rock
(shale) can be used for the correlation studies.
The most important molecules to provide such oil-oil and oil-rock correla-
tions with specificity are petroleum biomarkers, discussed in Sect. 5.6 of the last
chapter. Petroleum biomarkers, although present in trace amounts, are also important
molecules for in-depth geological and geochemical studies to relate the petroleum
with deposition environment and history. Their distribution patterns in crude oils
and source rocks unravel stratigraphic origin, migration pathway, accumulation, and
alteration due to biodegradation or water washing of the existing petroleum deposits
[7]. After all, petroleum is an organic mixture, so the biomarkers can serve as molec-
ular indicators for assessing oil and reservoir quality, especially in collaboration with
geophysical studies and measurements.
6.5 Hydrocarbon Migration 91
There are numerous factors that control the migration processes of hydrocarbons,
expansion of kerogens, increase of pressure, and the expulsion out of the source
rock. Kerogen generates a hydrocarbon fluids matrix at temperatures of at least
50–70 °C and a minimum depth range of 1000–1500 m. Overpressure creates fracture
porosity that allows the oil and gas to escape the impermeable source rock or to be
carried by pore water. The expulsion of the oil out of the source rock is also a dynamic
process driven by the oil generation itself. A good source rock has a total organic
content (TOC) ranging from 3 to 10%. The fluid pressure of the oil within black
shales can become high enough to produce microfractures in the rock. Once the
micro fractures form, the oil is squeezed out and the source rock collapses. Primary
migration of oil and gas is movement within the fine-grained portion of the mature
source rock. The released petroleum compounds can move through porous carrier
rock or faults into a trap. Hence, secondary migration is hydrocarbon movement
outside the source rock, or movement through fractures within the source rock,
shown in Fig. 6.5. Buoyancy of the hydrocarbons occurs because of differences
in densities of respective fluids and in response to differential pressures in carrier
and reservoir rocks. There is a possibility of tertiary migration when petroleum
moves from one trap to another or to a seep or even to the surface due to tectonic
movement, topographically driven flow, thermal expansion, etc [8]. The migration
mostly occurs as one or more separate hydrocarbons phases, liquid or gas, depending
on the temperature and pressure conditions. It’s also called dismigration [9].
As mentioned, the most essential elements of carrier and reservoir rocks are porosity
and permeability [10]. The permeable strata in an oil trap is known as the reservoir
rock. The reservoir rock must contain pores or voids to store hydrocarbon fluids,
shown in Fig. 6.6. These pores must be interconnected and permeable to fluids and
gases. The impermeable cap rocks or salt domes function as traps to stop petroleum
migration for accumulation. The majority of petroleum accumulation is found in
relatively coarse-grained porous and impermeable clastic rocks, such as sandstones
and siltstones, and carbonate rocks. Porosities in reservoir rocks usually range from
5 to 30%. Most of the traps where petroleum is found are formed by tectonic activity,
such as faults, folds, salt domes, etc. Fault zones are not always sealed as traps; they
might even provide migration avenues. Anticlinal structures are the most common
type of traps. Seals are provided by impermeable shales and claystones as cape rocks
shaped like inverted bowls, which often appear as hills on the surface. The famous
Spindletop well in East Texas was drilled into such a hill. There is always residual
water in pores which cannot be displaced by migrating petroleum.
92 6 Exploration for Discovery
Fig. 6.5 Schematic presentation of primary migration and secondary migration [10]
Fig. 6.7 Gas-oil and oil-water contacts in reservoir. OGC:oil-gas contact; OWC: oil-water contact
[3]
The process of the filling of reservoir traps and formation of petroleum accumula-
tions often occurs over extended periods of geologic time. At the Earth’s surface, the
traps are filled from the bottom with oil and gas due to gravity, since the petroleum
is lighter in density than the bottom water. The most common traps are the anti-
cline, where petroleum displaces the pore water from the top of the culmination and
expands into the flanks. The boundary where the oil-saturated and the water-saturated
pore spaces are in contact is referred to as oil/water contact.
Most reservoirs contain both gas and oil. Since the gas has the highest buoyancy,
a free gas phase will separate from the oil, accumulating in the apex of the structure
as a gas cap. If the source rock is “overmature”, more gas is generated than liquid and
the reservoir would be predominantly gas. Almost always, there is a large quantity
of water in porous rock under the hydrocarbon reservoir, as depicted in Fig. 6.7. The
gas above the liquid is under very high pressure. As such pressures, the solubilities
of gas-in-oil and oil-in-water are high. During production, safe operation requires
controlled degassing of the oil.
To hold the petroleum in the trap, an impermeable cap rock must seal the trap.
Some of the best sealing cap rocks are anhydrite or rock salt. The best quality cap
rocks hold many of the large petroleum accumulations are in the Middle East. Some
petroleum seeps occur where permeable pathways in the form of fractures or faults
lead to the surface of the Earth either from mature source rocks or from leaking
accumulations.
94 6 Exploration for Discovery
The vast Western Canadian Sedimentary Basin (WCSB) underlies 1,400,000 square
kilometres (540,000 square miles) of Western Canada, including southwestern Mani-
toba, southern Saskatchewan, Alberta, northeastern British Columbia, and the south-
west corner of the Northwest Territories (see the map in Fig. 6.8). It consists of a
massive wedge of sedimentary rock about 6 km (3.7 miles) thick under the Rocky
Mountains in the west, but the layer thins to zero at its eastern margins, shown in
Fig. 6.9 [11]. The WCSB contains one of the world’s largest reserves of natural gas,
estimated at 143 trillion ft3 with peak production 16 billion ft3 a day of gas in 2000.
It also contains the largest oil sand deposits, mainly in the areas of Athabasca, Cold
Lake and Peace River in Alberta, with an estimate of 200 million barrels of recover-
able oil. This compares to 298 million barrels of extra heavy oil in Venezuela and 268
million barrels of conventional oil in Saudi Arabia. [12] Conventional oil is being
depleted quickly. Oil sands accounted for ~90% of production from the WCSB by
2016. The WCSB also has huge reserves of coal.
The WCSB provides a great opportunity for geochemistry studies. The strata
were deposited at different ages, with various degrees of maturity in source rocks.
Bitumen on oil sands experienced severe weathering and biodegradation. Thus, the
oil, gas and bitumen provide an ideal set of geochemical samples for age, maturity
and alteration studies of marine-sourced oils and rocks.
Fig. 6.9 Cross section of Western Canadian Sedimentary Basin from southwest (left) to northeast
(right)
References
If the mud density is too low, a well is susceptible to a surface blowout. If the mud
density is too high, it can cause an underground blowout—the rupture of the reser-
voir underground—pushing drilling mud into another formation. The cuttings are
removed in shakers, and the fluid goes back to the mud pit, from which it is recycled.
A derrick is used to support drilling apparatus, which must be tall enough to
accommodate a 90-foot-long “triple” comprised of three sections of pipe. For off-
shore drilling, a platform must be built for support of the drilling rig. The drilling
system includes the power system (diesel engine and electric generator), the hoisting
system, rotating equipment, casing, the circulation system, and the blowout preven-
ters. The derrick, pulleys and hawser must be robust enough to support the lifting
and manipulation of the entire drill string, which can weigh hundreds of tons. The
kelly is the top joint of a drill string. It has flat sides that fit inside a bushing on the
7.1 Drilling for Oil and Gas Recovery 99
rotary table, which turns the drill string and bit. Note that not all drilling rigs use a
kelly system.
As the drill begins tunneling through the ground, chunks of rock and mud are
sucked into the pipe by pumps. The chunks are circulated out of the hole and into
pits on the surface previously carved by the drilling crew. The derrick is tall enough
to allow new sections of pipe to be added as the hole becomes deeper. After a certain
depth, depending on the oil reserve, the drill must be pulled out and the well must be
cased to prevent the hole from collapsing in on itself. Casing involves placing steel
piping into the hole and then pumping cement into the annular space between the
outside of the casing and the rock. Centralizers are sections of pipe which keep the
casing from resting against the surrounding wellbore wall, as in Fig. 7.2. Once this
section of the hole is deemed secure enough, the crew continues to drill deeper with
drill pipe of ever narrower diameters. This cycle of drilling, casing and cementing is
done until the final pre-determined depth is reached within the reservoir rock.
Throughout the drilling process, the location and direction of the drill bit is deter-
mined. Well logs are evaluated and cuttings are analyzed, and if necessary the drilling
plan is adjusted accordingly. When drilling is complete and tests have confirmed that
Fig. 7.2 Casing with bow-string centralizers that allow cement to flow through all empty annular
space between casing and well bore
100 7 Production for Recovery
the location is correct, preparations to extract the oil commence. A perforating gun
with explosive charges is dropped into the well to perforate the casing to allow the
flow of oil up the pipe.
Wells are completed by casing the well bore with steel pipe and cementing the
casing into place. Casing prevents the well from collapsing. The outside diameters
of casing pipe range from 4.5 to 16 in. (114–406 mm). Cementing is a key step. In
a good cement job, the entire annular space between the casing and the well bore is
filled with cement. Centralizers keep the casing from resting against the well bore to
block the flow of cement, resulting in a poor cement job and increasing the risk of a
blowout. Figure 7.3 shows well completion with cementing.
The integrity of cementing is tested by a positive-pressure test and a negative-
pressure test. In the positive-pressure test, the pressure in the steel casing is increased
to see if it is intact along with seal assembly. In the negative-pressure test, the pressure
is reduced to below atmospheric to test the integrity of the cement at the bottom of
the hole. The test is deemed successful if the pressure remains low after the suction
is stopped.
A key consideration during completion is the weight of the well casing, which rests
on the bottom of the well bore, creating friction that limits the distance/depth (D/D)
ratio. Long, shallow wells are especially susceptible to D/D limitations. In extended-
reach horizontal drilling (ERHD) technology, [2] a special tool at the bottom of
the casing allows the casing to be filled with air instead of mud. This reduces the
friction substantially, allowing much higher D/D ratios. The technology is employed
in certain wells at Wytch Farm, England. On Platform Irene offshore California,
ERHD enabled Unocal to set records, both for horizontal reach (14,671 ft), widest
pay zone (5990 ft) and greatest angle of deviation from vertical (76°) [3]. Those
records have since been broken, and the technology has advanced, but the invention
by Mueller, et al. [2] was a significant first step. After cementing, perforation guns
punch small holes through the casing into the reservoir rock, providing a path for the
flow of oil and gas into the well. In open-hole completion, the last section of the well
is uncased. Instead, the installation of a gravel pack stabilizes the casing and allows
fluids to enter the well at the bottom. After perforation, special acid-containing fluids
are pumped into the well to increase porosity and stimulate production. Usually, a
smaller diameter tube is inserted into the casing above the production zone and
packed into place. This provides an additional barrier to hydrocarbon leaks, raises
the velocity at which oil flows under a given pressure, and shields the outer casing
from corrosive well fluids.
A blowout preventer (BOP), shown in Fig. 7.4, seals the high-pressure drill lines
and relieves a sudden increase in well pressure into the atmosphere for uncontrolled
gush or gas. The BOP is a collection of safety valves and other devices at the top of
a well. It is located on the surface for an onshore well. It can be placed beneath the
ocean on the sea floor for an offshore well. When activated, it stops a blowout by
sealing off the top of the well.
Underground blowouts are the most common of all well control problems. Many
surface blowouts begin as underground blowouts. Prompt, correct reaction to an
underground event can prevent a dangerous and costly surface blowout [5]. The
severity of the BP Deepwater Horizon oil spill in 2010 was aggravated by the failure
of a blowout preventer.
102 7 Production for Recovery
Conventional vertical wells are drilled straight down into the earth. Descriptions of
such wells in China, drilled with bamboo poles, date back to 1500 BC. Vertically
drilled wells are only able to access the targeted gas and/or oil that immediately
below the well. Using horizontal or directional drilling techniques, shown in Fig. 7.5,
a number of wells can be drilled in different directions from a single well pad, which
is much more efficient than having numerous vertical well pads set up to extract oil
or gas. This decreases the surface disturbance and reduces the overall cost of well
pad setups, replacements and maintenance. This is especially beneficial offshore,
because it can decrease the required number of expensive platforms by an order of
magnitude. Tremendous savings are realized when offshore oil can be reached from
onshore drilling sites, as is the case for many wells in the huge Wytch Farm oil field
in the Purbeck District of Dorset, England.
Horizontal drilling technology achieved commercial viability during the late
1980s. Two key components in directional/horizontal drilling plays are mud motors
and measurement while drilling (MWD) sensors. Mud motors can rotate the drill
7.3 Directional Drilling/Horizontal Drilling 103
Fig. 7.5 Directional drilling (left) and horizontal drilling (right) wells [6]
bit without rotating the entire length of drill pipe between the bit and the surface.
Figure 7.6 shows bent sections of pipe (bent subs) which compel the bit to follow a
path that deviates from the previous orientation. MWD sensors are used determine
the azimuth and orientation of the bit. A rotary steerable system (RSS) employs spe-
cialized downhole equipment to replace conventional directional tools such as mud
motors, allowing operators to steer the drill bit in real time.
In the past few years, directional drilling combined with hydraulic fracturing has
been applied to tight rock formations, resulting fantastic production of oil and/or
natural gas from reservoirs which were otherwise unproductive. Major examples are
the Eagle Ford Formation near Three Rivers, Texas, the Bakken Formation of North
Dakota, and the Marcellus Shale of the Appalachian Basin. Due to this technology,
the United States is now among the top oil producers in the world.
The availability of directional/horizontal drilling also stimulates the development
of in situ toe-to-heel air injection (THAI) combustion and steam-assisted gravity
drainage (SAGD) production of heavy oil and bitumen to be discussed later.
The term of “offshore drilling” refers to drilling activities on the continental shelf,
although the term can also be applied to drilling in lakes, inshore waters and inland
seas. In offshore drilling, a well is drilled below the seabed to explore for and sub-
104 7 Production for Recovery
Steering tool
Orienting sub
Bent sub or
bent mud
motor housing
Bit
Fig. 7.6 Arrangement of a steering tool, orienting sub, and bent sub for directional drilling
sequently extract petroleum which lies in undersea rock formations. Offshore wells
can be drilled from onshore or from platforms.
For offshore wells, two depths are important: the depth of the water, and the
distance from the top of the sea bed to the pay zone [7]. Both must be considered when
selecting a platform design. Another consideration is metocean, i.e., meteorology
and oceanography, involving the quantification of winds, waves, currents and related
physical phenomena in the ocean and atmosphere. United Nations Convention on
Law of the Sea is used for the determination how far we can drill offshore from the
coastline. Figure 7.7 shows the main offshore oil production region in the world as
of 2012 [7].
7.4 Offshore Drilling 105
Fig. 7.7 Main offshore oil production regions in the world as of 2012 [7]
raise the rig above the surface and keep it safe from choppy waters; an example is
shown in Fig. 7.8. Semisubmersible rigs and drill ships are used in deeper waters.
Semisubmersible rigs simply let in enough water to lower the platform to appropriate
operating heights. The weight of the lower hull stabilizes the drilling platform, while
massive anchors hold it in place. Drill ships that have a drilling rig on the top deck use
dynamic positioning equipment to maintain alignment with the drilling site, guided
by satellite information and underwater sensors on the subsea drilling template to
keep track of the drilling location.
The major production structures and systems are summarized in Fig. 7.9 [4]. Fixed
platforms or jackets are used for water less than 1500 ft deep with mild climates.
Locations suitable for such platforms are not only constrained by water depth, but
also by the wind and wave strength. Compliant towers are designed to sway and
move with the stresses of wind and sea. Sea Star platforms are a larger version of
submersible designs and are connected to the ocean floor by tension legs. Both of
these platforms operate at water depths up to 3500 ft. For deeper waters, floating
production systems, tension leg platforms (TLP), and subsea systems transfer the
oil and natural gas to production facilities, either by risers or undersea pipelines.
For waters deeper than 7000 ft, a spar platform on a giant, hollow cylindrical hull is
used. The most sophisticated and versatile floating production systems are floating,
production, storage and offloading (FPSO) platforms. FPSOs include an offloading
and storage capacity, They are the obvious choice for stranded fields where there is
no existing pipeline infrastructure. FPSOs store oil, typically about a million barrels,
and offload that oil to a tanker. They are typically fitted with extensive onboard
separation and processing equipment. An example of FPSOs is shown in Fig. 7.10.
As mentioned earlier, the greatest water depth a jackup can drill in is 550 ft.
Many newer jackup units have a rated drilling depth of 35,000 ft. On the floating rig
side, the deepest water depth so far is 12,000 ft. A handful of these rigs have a rated
drilling depth of 50,000 ft, but most of the newer units are rated at 40,000 ft [9].
In primary oil recovery, fluids are pushed into the production well and up to the
surface by natural forces: gas drive and water drive. Gas drive is the most efficient.
Oil is pushed by natural underground pressure, usually supplied by associated natural
gas, liquid expansion and evolution of dissolved gas, shown in Fig. 7.11. Once the
wellbore is drilled, the free gas begins to expand. The expansion energy of the gas
is what allows it to rise out of the reservoir and travel to the surface. The gas-oil
contact plane drops as the oil is depleted. The next most efficient propulsive force
is natural water drive, in which the oil is driven upward under hydrostatic pressure
and into the well by expansion of water inside the reservoir, also shown in Fig. 7.11.
Below the natural resource is an aquifer. The water that drives this kind of recovery
can be either located beneath the natural resource or on the edges of the reservoir in
which the resource is contained. Once the wellbore is drilled into the reservoir, water
7.5 Primary Recovery—Primary Oil Production 107
in the aquifer begins to push the hydrocarbons to the surface until they have been
completely displaced or until the point at which so much water has accumulated in
the well that it is no longer a quality resource.
The efficiency for primary recovery is generally low. Primary recovery typically
recovers 5–15% of oil in the reservoir due mostly to microscopic trapping and bypass-
ing of the remaining oil [10].
108 7 Production for Recovery
Fig. 7.10 A colossal offshore platform lights up the night off the coast of Norway [8]
Fig. 7.11 Primary oil recoveries by gas drive and water drive
7.6 Secondary Recovery 109
In primary recovery, the natural force used to drive the resource to the surface will
eventually decrease and become not sufficient. To compensate, an artificial lift system
using mechanical energy may be implemented to aid in the recovery of the natural
resource in a more economical fashion. This mechanism is secondary recovery.
Horsehead pumps or sucker-rod pumps are common surface implements. Submerged
pumps also are used. As with primary recovery, a single production well can be used
for artificial lift.
Water injection (water flood) or gas injection (gas flood) stimulate production by
increasing reservoir pressure or displacement of oil towards the production wells.
At least two wells are used: an injection well and one or more production wells.
Typically, the gas is reinjected natural gas. Another method is gas lift in which
compressed air, water vapor, carbon dioxide or some other gas is injected into the
bottom of an active well, reducing the overall density of fluid in the wellbore. Gas
injection delivers a necessary amount of compressed gas to a distribution cap, pushing
the oil out of the reservoir. Water injection involves pumping water into the reservoir
to displace the natural resource to an adjacent production well.
Figure 7.12 generically illustrates the use of a combination of injection and pro-
duction wells to stimulate production, via both secondary and tertiary recovery. On
average, primary and secondary recovery methods combined allow 20–40% of the
reservoir oils to be recovered [11].
Fig. 7.12 Generic depiction of injection and production wells for secondary and tertiary recovery
110 7 Production for Recovery
Tertiary recovery, also known as enhanced oil recovery (EOR), is used to extract
the remaining oil from reservoirs where primary and secondary recoveries are no
longer cost effective. Many sandstone or carbonate reservoirs have low primary and
secondary recovery due to poor sweep efficiency for bypassed or unswept oil. EOR
may also be used to stimulate production from reservoirs containing very viscous
crude oils and from low-permeability carbonate reservoirs. It is designed to reduce
viscosity of the crude oil. According to the US Department of Energy (DOE), there
are three primary techniques for EOR: gas injection, chemical injection and thermal
recovery. Using EOR, 30–60%, or more, of the formation and reservoir’s original
oil can be extracted, compared with 20–40% using primary and secondary recovery.
Miscible flooding is considered one of the most effective enhanced oil recovery
processes applicable to light-to-medium oil reservoirs. This method can yield as
much as 17% of a field’s original oil-in-place. It is accomplished with hydrocarbon
solvents or gases such as carbon dioxide, natural gas, or liquefied petroleum gas
(LPG). The injected fluids are capable of displacing crude oil for recovery from the
reservoir rock. Supercritical CO2 has a viscosity similar to hydrocarbon miscible
solvents. Both improve volumetric sweep-out when there is an unfavorable viscosity
ratio in the reservoir. However, the CO2 density is similar to that of oil. Therefore,
CO2 floods minimize gravity segregation compared with the hydrocarbon solvents.
Introducing miscible gases reduces the interfacial tension between oil and water,
maintains reservoir pressure, and improves oil displacement. CO2 is most commonly
used, because it reduces the oil viscosity and is less expensive than LPG.
One difficulty with CO2 injection is that petroleum companies must first secure,
transport, and store an adequate supply of CO2. However, CO2 is a green-house gas
produced in large volumes by many industrial factories. CO2 capture and sequestra-
tion (CCS) by using it to stimulate oil production mitigates two important problems.
Certain producers inject CO2 injection into coal mines to produce coal-seam methane
[12].
Not all locations and reservoirs are appropriate for this technique; Suitability
depends on geology and fluid characteristics. Nitrogen gas can be used in combination
with CO2 flooding when complete CO2 flooding is not economical. Nitrogen and
flue gas are lower in cost and have shown success in re-pressuring reservoirs.
7.7 Tertiary Recovery (Enhanced Oil Recovery) 111
There are many chemicals that can improve oil mobility by reducing its viscosity
and/or reducing the intermolecular interactions which hold oil onto rock. Chemical
methods include polymer flooding, surfactant flooding and alkaline flooding. In all
of the chemical injection methods, the chemicals are injected into several wells
(injection wells) and the production occurs in other nearby wells (production wells).
and the organic acids naturally occurring in the oil. It enhances oil recovery by
lowering the interfacial tension, decreasing the rock wettability, emulsification of
the oil, mobilization of the oil, and helping to draw the oil out of the rock.
Caustic flooding is usually accompanied in conjunction with surfactant and poly-
mer flooding. The combination is called alkaline-surfactant-polymer (ASP) flooding,
shown in Fig. 7.13, in which the three slugs are used in sequence. Alternatively, the
three fluids could be mixed together and injected as a single slug. The objective of
the ASP flooding process is to reduce the chemical consumption per unit volume of
oil, resulting in a reduction in cost.
Microbial injection is part of microbial enhanced oil recovery (MEOR) and is rarely
used because of its higher cost and because the method is not widely accepted.
These microbes function either by partially digesting long hydrocarbon molecules,
by generating biosurfactants, or by emitting carbon dioxide which can be used in gas
injection.
7.8 Thermal Recovery 113
oil + water
steam
Various methods are used to heat the crude oil in the formation to reduce its viscosity
and surface tension, thus, increasing the permeability of the oil. The heated oil may
also vaporize and then condense forming improved oil. Methods include cyclic steam
injection (Huff and Puff), steam flooding, and combustion.
Cyclic steam stimulation (CSS) is also known as the Huff and Puff method, shown
in Fig. 7.14. It requires only one wellbore and consists of 3 stages: injection (Huff),
soaking, and production (Puff). First, steam at elevated pressure is injected into a
well at a temperature of 300–340 °C for a period of weeks to months. Next, the
well is allowed to sit for days to weeks to allow heat from the elevated pressure
steam to soak into the formation and reduce the viscosity of the oil around the well.
Finally, the hot oil is pumped out of the well for a period of weeks or months.
Once the production rate falls off, the well is put through another cycle of injection,
soaking, and production. The process is used for thinner shallow production near
bitumen reservoirs. High reservoir porosity and oil saturation is ideal for this method
of recovery. This process is repeated until the cost of injecting steam becomes higher
than the money made from producing oil [13]. This process can typically remove
about 25% of the total reserve.
Steam degrades (weakens) the formation by removing alkali- and alkaline-earth
ions from reservoir constituents, such as particulates, thereby causing the forma-
tion to swell. This decreases permeability and slows down production. Watkins and
Kalfayian [15] invented methods for decreasing swelling by injecting ammonium
salts and ammonia precursors with the steam, replacing the removed K+ , Mg2+ , etc.,
with NH4 + .
114 7 Production for Recovery
Steam flooding is also known as steam drive injection or steam stimulation, shown
in Fig. 7.15. In this method, some wells are used for steam injection and other
wells for oil recovery. Two steps are involved: first, the oil is heated by steam to
higher temperatures to decrease its viscosity so that it more easily flows through the
formation toward the producing wells; second, the oil is pushed to the production
wells in a similar manner as water flooding. More steam is needed for this method
than for the cyclic method.
The steam injection methods mentioned above are applicable for oils with viscosities
in the range of 100–10,000 cP. Heat provided by the steam reduces the viscosity of
the oil, thereby making it mobile. For heavy oils and bitumen with viscosities greater
than 10,000 cP, combustion becomes the method of choice.
An effective and revolutionary in situ combustion method for producing heavy
oil is THAI (toe-to-heel air injection) [16], shown in Fig. 7.16. THAI combines a
Fig. 7.15 Steam drive injection (steam flooding or steam stimulation) [12]
7.8 Thermal Recovery 115
vertical air injection well with a horizontal production well. THAI is also applicable
to horizontal wells, which due to their geometry are not suitable for steam injection
methods.
For the first few months, steam is injected in the vertical well to preheat the hori-
zontal well and condition the reservoir around the vertical well. Then air is injected
in the vertical well and combustion initiated. The combustion raises temperatures
to approximately 400–600 °C (750–1110 °F). At these temperatures, both thermal
cracking and coking occurs. In this process, about 10% of the oil is lost to coke. The
oil from thermal cracking is of higher quality than reservoir oil. The mobilized oil
flows by gravity as a toe to the horizontal section of the L-shape production well.
The combustion front sweeps the oil from the toe to the heel of the horizontal pro-
ducing well, recovering the original oil-in-place while partially upgrading the crude
oil in situ. The combustion gasses bring the mobilized oil and vaporized water to the
surface, so no pumps are needed.
Another good feature of this process is the minimal amount of water supply
needed. Once the first few months of steam injection has been completed, no more
water or even natural gas is used. Combustion continues as long as air is injected.
With less equipment also comes a smaller footprint at the surface and less land is
required. The combustion is also self-limiting, so as soon as air injection is stopped
the flames burn out. With this method, the area at and near the combustion zone will
turn into coke Eventually, the reservoir no longer retains any oil or natural gas.
Fig. 7.16 Toe-to-heel air injection (THAI) for oil recovery [17]
116 7 Production for Recovery
Heavy crude oil or extra heavy crude oil is highly viscous, and cannot easily flow
under normal reservoir conditions. Heavy crude oil has been defined as any liquid
petroleum with an API gravity less than 20°. In 2010, the World Energy Council
defined extra heavy oil as crude oil having a gravity of less than 10° and a reservoir
viscosity of no more than 10,000 cP. Compared to the lighter crude oils, heavy
crude oils have higher viscosity and specific gravity, as well as heavier molecular
composition with significant contents of nitrogen, oxygen, and sulfur compounds
and heavy-metal contaminants.
Natural bitumen, including tar sands or oil sands, shares the attributes of heavy
oil but is yet more dense and viscous, having a viscosity greater than 10,000 cP.
Natural bitumen and heavy oil resemble the resids from the refining of light crude
oil. They are thought to be the residue of formerly light oil that has lost its light-
molecular-weight components through degradation by bacteria, water-washing, and
evaporation (weathering). Conventional heavy oils and bitumens differ in the degree
by which they have been degraded from the original crude oil.
According to World Resources Institute, remarkable quantities of heavy oil and
oil sands are found in Canada and Venezuela. The largest reserves of heavy crude
oil in the world were located north of the Orinoco Basin in eastern Venezuela. It was
estimated that there were 270 billion barrels of recoverable heavy or extra-heavy
oil reserves in the area, similar to the amount of conventional oil reserves in Saudi
Arabia.
Heavy oil and bitumen are recovered in several ways. They can be dug out with
conventional mining techniques, or they can be liquefied by the injection of high-
pressure steam, for example in cyclic “huff and puff” operations (see above).
At surface facilities, notably in Venezuela and Canada, recovered heavy oil and
bitumen are diluted with lighter hydrocarbons. The resulting “Dilbit” (diluted bitu-
men) flows under ambient conditions, so it can be transported conventionally in
pipelines and oil tankers.
Kerogen is recovered from oil shale by several methods. From 1985 to 1990,
Unocal recovered some 4.6 million barrels of synthetic crude oil from oil shale in a
complex mining and upgrading venture at Parachute Creek, Colorado [18]. The plant
yielded roughly 40 gallons of oil per ton of rock. In the vertical-shaft retort, crushed
kerogen-containing shale was pumped up from the bottom of the retort vessel. Hot
recycle gas flowed counter-currently downward, decomposing the rock and releasing
hydrocarbons. Condensed shale oil was removed from the retort at the bottom. Part
of the hot gas was recycled. The rest either was used to produce heat and hydrogen,
or recovered as product. The spent shale was removed from the top of the retort,
cooled, and stored in pits or returned to the mine. In the reducing environment of the
retort, sulfur and nitrogen were converted to H2 S and NH3 , which were recovered
from product gases by conventional means. The plant yielded high-quality synthetic
crude oil suitable for further refining in conventional facilities.
7.9 Recovery of Heavy Oils and Bitumen 117
More discussions on diluted bitumen and synthetic crude oil for transportation
can be found in Chap. 17, which covers mid-stream operations.
Other oil shale processes involve partial combustion, either underground, at the
surface, or in shafts drilled horizontally into kerogen-rich formations. From 1972
to 1991, Occidental Petroleum developed an in situ process, in which explosives
were used to create underground chambers of fractured oil shale. The oil shale was
ignited with external fuel, and air and steam were injected to control combustion.
The hot rock fractured and released shale oil, which was pumped to the surface from
a separation sump and collecting well.
Bitumens derived from oil shale and many tar sands contain small but significant
amounts of arsenic, which are severe poisons for catalysts in refineries.
Oil sands are recovered using two main methods: open-pit mining and in situ drilling.
The method depends on how deep the reserves are deposited. In Alberta, 97% of the
total surface area of the oil sands region could be developed in situ.
Approximately 20% of the oil sands lie close enough to the earth’s surface to be
mined, which impacts 3% of the surface area of the oil sands region.
Open-pit mining is similar to many coal-mining operations. Large shovels scoop
the oil sands into trucks, which take it to crushers, where the large clumps are broken
down. The oil sand is then mixed with water and transported by pipeline to a separa-
tion plant. A combination of very hot water, agitation, and other processing methods
are required to increase bitumen separation to about 75%. The resulting bitumen is
118 7 Production for Recovery
not very usable, though, and has to be upgraded. Roughly two tons of sand must
be processed to produce just one barrel of useable oil, so the return on investment
for this method is extremely low; not to mention all of the negative environmental
impacts. Compared to other methods, it has the largest land-use footprint, produces
the most greenhouse gases. It negatively affects surrounding wildlife, and pollutes a
vast amount of water. So companies generally try to avoid this if possible.
Tailings ponds are an operating facility common to all types of surface mining.
In open-pit oil sands processing, tailings consisting of water, sand, clay and residual
oil are pumped to these ponds, where settling occurs. According to industry sources,
78–86% of the water is recovered and reused [20]. Supposedly, when the ponds are
7.9 Recovery of Heavy Oils and Bitumen 119
One of the simpler methods of heavy oil extraction is the Cold Heavy Oil Production
with Sand method or CHOPS. This method is simpler because it does not require
sand filtration, allowing sand and oil to be extracted together. The method works
because removing sand creates more space within the reservoir, allowing larger
pockets of liquid oil to form. It’s called “Cold” because there no heat is injected into
the reservoir, so it’s also relatively energy efficient. This method isn’t super effective
and can only recover about 5–6% of the oil within the deposit, but since there is
no heat injection it is pretty cheap to employ. The disposition of the oily sand, after
it has been separated from the extracted liquid, is challenging. Some is used in the
construction of asphalt roads, but there are some problems with that too. Most is
currently deposited in underground salt caverns.
The majority of the oil sands lie more than 70 m (200 ft) below the ground and are
too deep to be mined. These reserves can be recovered through wells with thermal
stimulation, as discussed above.
A form of steam flooding that has become popular since 1996 in the Alberta
oil sands is steam assisted gravity drainage (SAGD), in which two horizontal wells
are drilled, one a few meters above the other. Steam is injected into the upper well
as shown Fig. 7.18. The steam injection can also be performed by multiple verti-
cal injection wells above the horizontal production well. The steam increases the
temperature of the oil sand around it. The viscosity of the oil sand decreases as the
temperature increases. The oil sand and condensed steam trickle down due to gravity.
The condensed steam and less viscous bitumen from the oil sands flow into the lower
horizontal production well. Pressure pushes the bitumen and water into the storage
tanks above ground. The bitumen-water mixture is separated into bitumen and water.
The bitumen goes off to be refined while part of the water is recovered and recycled
within the process. SAGD is a unique process that has taken precedence, due to its
higher recovery, lower cost, and greater efficiency. The development of SAGD is
fairly new in comparison to other methods of oil recovery. The progress of drilling
techniques overlapped with the development of SAGD, so the use of horizontal wells
became less expensive and more efficient. SAGD recovery ranges from 50 to 70%
of oil-in-place.
In SAGD, some of the lighter contaminants will rise with the steam and won’t
have to be separated later. This doesn’t happen with a lot of other heavy oil extraction
methods. However, it has some drawbacks, one being just the cost of generating the
120 7 Production for Recovery
large amount of steam required. Variations on this method employ solvents instead
of steam to increase the liquidity of the oil, making the process more energy efficient.
Another drawback is high water use. There is a concern that it relies too much on
nearby water supplies and will deplete lakes and river streams. There have been
instances of contamination due to leaking wells. Treating the contaminated water is
costly and not always done [22].
Natural gas is available as associated gas in crude oil reservoirs and as non-associated
gas in natural gas reservoirs and in the form of condensates. Recently, unconven-
tional sources, such as shale gas, tight gas and coal-bed methane, are commercially
exploited at large scales [24]. Unconventional resources are defined as those that can-
not be produced commercially without altering rock permeability or fluid viscosity.
7.10 Hydraulic Fracturing (Fracking) 121
Tight gas refers to natural gas reservoirs produced from reservoir rocks with
such low permeability, having less than 0.1 millidarcy (mD) matrix permeability and
less than 10% matrix porosity, that considerable hydraulic fracturing is required to
harvest the well at economic rates. These reservoirs do not have depth constraints.
They can be developed deep or shallow, at high or low pressure and temperature, in
stacked multilayered or single layer configurations, and in homogeneous or naturally
fractured formations.
It has been estimated that shale contains as much as 30% of today’s oil reserves.
Oil and gas are trapped in fine fissures and cannot be recovered conventionally.
Hydraulic fracturing (Fig. 7.19), a well stimulation technique also known as fracking,
is a process used in nine out of 10 natural gas wells in the United States. High quality
crude oils are also produced. Fracking involves injecting a specially designed fluid
under controlled pressure intermittently over a short period (three to five days) to
create fractures in a targeted deep-rock formation. The fractures permit oil, natural
gas and brine to flow to the wellbore.
A well drilled for hydro-fracking goes vertically down until it hits shale. It then
is drilled horizontally through the shale for up to 5000 ft. The next step is injection
of fluid, which has a wide variety of potential compositions. The “fracking fluid” is
typically a slurry of water (90%), proppant (9.5%) and chemical additives, such as
122 7 Production for Recovery
thickening agents. gels, foams, light diesel, compressed gasses, or any of the other
approximately 750 chemical additives registered for this purpose. The goal is to
increase permeability of the surrounding rock to allow any oil released to flow more
easily. The proppants are small grains of sand, treated like resin-coated sand, man-
made ceramic beads, aluminum oxide or other particulates used to hold the fractures
open when the hydraulic pressure is removed from the well to recover gas and oil.
The actual fracturing part, though, is caused by extremely high pressure and
velocity, sometimes as much as 15,000 psi and 265 L/min. The resulting cracks free
fluid hydrocarbons, including gases, light oils, and heavy oils, allowing them to flow
more easily into the well. Typically, the additives decrease surface tension, reduce
viscosity, and prevent emulsions.
The injected fluids are somewhat recovered with the oil and disposed by injection
into deep wells. This so-called “frack water” is highly contaminated with dissolved
minerals, toxic trace metals, and heavy oil. It could be recovered, but only at great
expense. The disposition of frack water is a question which causes great public
concern.
To check on the status of fracturing and to possibly catch any leakage before it gets
out of hand, seismic monitoring is used. The length and depth of the fractures can be
measured and compared with expectations. If problems are discovered, especially
those that may cause environmental problems, fracturing can be stopped or at least
reduced. Monitoring also can reveal close-by reservoirs. After fracturing has been
completed, the frack oil is handled the same way as conventional oil.
Hydraulic fracturing is highly controversial due to the potential for adverse envi-
ronmental impact. Responsible operators apply fracking well below water tables and
cement their wells using best practices. When such practices are applied, the like-
lihood of loss of containment is minimal. However, disreputable companies have
fractured with explosives instead hydraulic pressure and have done so at shallow
depths, where problems are more likely.
Fracking affects the quality of air due to the amount of fuel being burned to run
the pumps at such high pressure. Some reports also say that up to 90% of the fractur-
ing liquid remains underground and cannot be reclaimed; over time, this may affect
nearby water resources. In some areas, fracking increases seismic activity substan-
tially. Prior to injection into disposal wells, wastewater is stored at the surface in open
vats, from which degassing contaminates the air with methane and other greenhouse
gases. The U.S. congress exempted some aspects of fracking from the Safe Drinking
Water Act (SDWA). The EPA has strict regulations on what oil companies can do
at their fracturing sites, but nobody can know in advance exactly what may happen
underground, where most of the problems can occur.
The oil and gas that cannot be recovered by conventional means, such as primary
to tertiary recoveries, are classified as unconventional oil and gas (UCOG). UCOG
includes oil sand, bitumen, extra-heavy oils, gas hydrates, coalbed methane, gas in
tight sand, etc. The oil and gas recovered by hydraulic fracturing are known as shale
oil and shale gas. The shale oil from fracking can be confused with the oil obtained
from retorting oil shale. Shale oils from fracking are high quality oils because the
heavy metals and asphaltenes remain in the reservoir or formation. But shale oils from
7.10 Hydraulic Fracturing (Fracking) 123
oil shale retorting contain the heavy fractions and toxic inorganic materials, such as
arsenic and mercury. It would probably be better to refer the shale oil from fracking
as “fracking oil” to differentiate from the “shale oil” from oil shale retorting. Shale
gas from fracking is natural gas, mainly methane. The gas from oil shale retorting
contains methane, too, but it also contains significant C2 -to-C4 gases.
A primary consideration when extracting unconventional petroleum, such as from
SAGD and hydraulic fracturing, is the net-energy gain, or NEG. The NEG of a process
is the amount of energy available in the oil recovered minus the amount of energy
that had to be put into retrieve it. If this value is not high enough then it is not worth
recovering the petroleum.
Crude oil comes from the ground mixed with a variety of substances: gases, water,
salt, and dirt. These must be removed before the crude can be transported effectively
and refined without undue fouling and corrosion. Some cleanup occurs in oil fields
and midstream processes such as the preparation of syncrudes. Natural gas is mostly
methane, but it may contain hydrogen sulfide, CO2 , water, higher hydrocarbons,
mercury compounds, and noble gases (He and Ar).
Natural gas and crude oil need to be transported to a gas plant or a refinery, usually
a considerable distance away from the fields, for processing into useful products.
For large-scale transportation, pipelines and tankers are commonly used. For smaller
scales, especially for the distribution of petroleum products, railroad cars, barges and
tank trucks are used to a large extent. Prior to transportation, most gas and oil require
some form of pretreatment near the reservoir to meet transportation requirements
and safety specifications/regulations.
Considerable planning, including possible trading, is involved in determining how
and where to ship to the oil and gas. Transportation and trading, even storage, can be
considered as “midstream” operations between the field production (upstream) and
refinery or gas plant (downstream).
In certain cases, pretreatment of natural gas can be conveniently performed at the
wellhead. The produced oil, on the other hand, is usually collected from several wells
and sent to a central facility for separation of gas, oil, water and sand.
A field separator at well site is often no more than a large covered vessel that
provides enough residence time for gravity separation into four phases: gases, crude
oil, water, and solids. Generally, the crude oil floats on the water. The water is
withdrawn from the bottom and is disposed of at the well site. Gases are withdrawn
from the top and piped to a natural-gas processing plant or reinjected into the reservoir
to maintain well pressure. Crude oil is pumped either to a refinery through a pipeline
or to storage to await transportation by other means.
Low density natural gas is mostly moved by pipelines. It is also transported by
seagoing tankers at ambient or atmospheric pressure, with the cargo under refriger-
ation; special alloys to resist brittleness at temperatures as low as −160 °C became
124 7 Production for Recovery
available in the late 1960s, enabling operators to cryogenically liquefy natural gas
(LNG) for shipping.
More discussions in transportation can be found in Chap. 17.
References
Bitumen was used in ancient times for many reasons—for water-proofing boats
and canals, as mortar for bricks, as a medicinal ointment, for making spears and
other weapons, and in mummification. Liquid petroleum and natural gas were used
mainly for heating, lighting, cooking, and warfare. Ancient writers referred to flaming
arrows and firepots based on pitch, turpentine, naphtha, sulfur, and charcoal. How the
naphtha was obtained is a subject of debate, but by 300 AD, it was being recovered
with simple distillation equipment such as that described by the 3rd century Greek
alchemist Zosimos of Panopolis, shown in Fig. 8.1. Only small amounts of light
fractions were recovered from petroleum for separate use. The ancient people wanted
bitumen, mostly, so they allowed petroleum to “age” by storing it in open pits,
allowing the light ends evaporate.
Naphtha was an essential component of Greek fire, which was invented during the
reign of Constantine IV Pogonatus (668–685) by Callinicus of Heliopolis, a Jewish
refugee from Syria. The composition of Greek fire is still debated, but suggestions
include combinations pitch, naphtha, quicklime, sulfur, and saltpeter. The substance
could be thrown in pots or discharged thorough syphon tubes. It ignited on contact
with air and could not be extinguished with water. In 673, Greek ships defended
Constantinople with this formidable weapon, crippling an Arab fleet that was trying
to attack the city.
The demand for petroleum started increasing after the invention of kerosene lamps
in the mid- and late-19th century. Since then, several historical events influenced
the evolution of today’s sophisticated technology for refining petroleum into useful
fractions. The progress in development is essentially market driven. The market is
driven, of course, by supply and demand. From about 1970 to 2014, the Organization
of Petroleum Exporting Countries (OPEC) largely succeeded in controlling prices
by controlling supply. Starting in 2011, OPEC’s influence diminished due to oil
production by hydraulic fracturing in the United States, which dropped the spot
price of bench-mark crudes from >$110 per barrel in mid-2014 to <$30 per barrel
Fig. 8.1 Distillation equipment described by 3rd Century Greek alchemist Zosimos of Panopolis
[1]
in early 2016 (in 18 months). Since then, the price gradually recovered to above $70
per barrel today (3Q, 2018). Figure 8.2 shows how oil prices have varied since 1859,
and how prices have corresponded to certain world events.
Before they can be used, kerosene, gasoline, and other fractions must be separated
from crude petroleum or other sources in a refinery. The earliest known refinery was
built in 1745, at the behest of the Empress Elisabeth of Russia. The plant was built
near an oil well in Ukhta by Fiodor Priadunov to distill a kerosene-like substance
from “rock oil” for use in oil lamps by Russian churches and monasteries.
For illumination, kerosene was superior to vegetable oils and most animal fats.
Kerosene burned cleanly and steadily, while most animal fats tended to burn with
smoky flames. The best animal fats for lamps came from whales. In the 1850s, the
over-hunting of whales was threatening the survival of several species. Prices were
fluctuating wildly. Experts agree that petroleum saved the whales [3].
In 1847, James Oakes built a “rock oil” refinery in Jacksdale, England. The unit
processed 300 gallons per day to make “paraffin oil” for lamps. Around the same
time, James Young built a coal-oil refinery in Whitburn, Scotland.
Samuel M. Kier of Pennsylvania is credited to give birth to the U.S. refining
industry. Kier recovered crude oil as a by-product of his brine well, which was the
basis of his salt business. Originally, he sold the crude oil as a medicine. In 1849,
8.1 Significant Events in Petroleum Processing 131
Fig. 8.2 Selected world events and crude oil prices, 1859–2015. Prices are annual averages. Items
in purple boxes indicate the start dates for major new production. The underlying format comes
from the BP Statistical Review of World Energy—2015. Recent prices come from the US Energy
Information Administration [2]
when the medicine fad faded, he successful refined the liquid petroleum to produce
“Carbon Oil” for kerosene lamps and started building the first U.S. petroleum refinery
in Pittsburgh, Pennsylvania.
The market for kerosene boomed—supply drove demand, in this case—and the
demand for whale oil plummeted after Edwin L. Drake struck oil near Titusville,
Pennsylvania in 1859, triggering the Pennsylvania Oil Boom. By the end of the
1860s, there were 58 refineries operating in Pittsburgh alone.
In 1870, John Rockefeller and five partners established the Standard Oil Company
and began consolidating the U.S. refining business. Less than 10 years later, Standard
Oil controlled 90% of US refineries and pipelines. Critics accused Rockefeller of
engaging in unethical practices, such as predatory pricing and colluding with railroads
to eliminate his competitors, in order to gain a monopoly in the industry. In 1911, the
U.S. Supreme Court ruled Standard Oil to be in violation of anti-trust laws and ordered
it to dissolve. Although Rockefeller is vilified as the ultimate “robber baron,” he
probably saved the industry from itself. In other parts of the world, notably Austrian
Galicia, unrestrained over-production, even when prices were very low, irreparably
depleted one of the world’s largest oil fields and damaged the nearby environment;
evidence of that damage lingers even today, more than 100 years later [4].
The use of kerosene lamps started to decline after the invention of light bulbs by
Thomas Edison in 1878. At the same time, several internal combustion engines were
132 8 Petroleum Processing and Refineries
invented. In 1889, Gottlieb Daimler, Wilhelm Mayback and separately Karl Benz
built gasoline powered automobiles. Rudolf Diesel is credited for building a high-
compression prototype engine in 1897 after Akroyd Stuart built the first working
diesel engine in 1892. Such engines burned fuels that are difficult to vaporize in a
gasoline engine. A major development was the switch from coal to oil in warships.
In the 1890s, Germany was talking about fueling new ships with petroleum. In the
1910s, Winston Churchill, in his role as First Lord of the Admiralty, pushed the Royal
Navy in this direction. With coal, he explained, storage bunkers closest to the boilers
were emptied first. As the coal inventory decreased, the distance between the boilers
and the supply increased, so it took more time and/or more sailors to bring coal via
conveyors to the boilers. This could handicap a ship during the heat of battle, when
the need for both steam power and manpower were highest. Liquid petroleum could
be pumped, so inventory didn’t affect the availability of fuel or combatants—unless
of course the supply ran out. Britain’s need to fuel its navy far from home led to
the formation of the Anglo-Persian Oil Company, the precursor to British Petroleum
(today’s BP).
The invention of internal combustion engines shifted the main market driver for
petroleum from illumination to transportation. The revolutionary moving assembly
line production of the Model T by Henry Ford in 1908 made automobiles affordable to
the general public. Demand for gasoline jumped, and continued to rise steadily—with
a pause during the Great Depression—until the 1970s, when unrest in the Middle
East triggered events that dampened supply.
Steady improvements in design and manufacturing technology have decreased
transportation costs and increased the maximum size of crude distillation units. Unit
operating costs are lower in large industrial areas (zones) such as Rotterdam; Singa-
pore; Chiba/Yokohama; South Korea; Sarnia, Ontario; Greater Los Angeles; Greater
Chicago; Greater Seattle; the San Francisco Bay Area; and largest of all, the U.S. Gulf
Coast between Corpus Christi, Texas and Mobile, Alabama. Through the Colonial
and Plantation Pipelines, the Gulf Coast Basin supplies a significant percentage of the
fuel consumed in the Northeast United States. In basins, several refineries and petro-
chemical plants can share offsites (power, steam, industrial gases) and have ready
access to nearby suppliers of chemicals and catalysts, engineer service companies,
not to mention a large, concentrated pool of trained personnel.
These improvements in technology, coupled with basin economics, led to con-
siderable consolidation—the merging of two or more adjacent facilities—and to the
shutdown of smaller, isolated, less efficient plants. From 1981 to 2014, the number of
operating U.S. refineries went from 321 to 123. At the same time, the average U.S.
refinery capacity rose from about 55,000 barrels per stream day (BPSD) to about
146,500 BPSD.
Table 8.1 summarizes the significant events in petroleum processing from 1860
to 2000 [5].
8.1 Significant Events in Petroleum Processing 133
Fig. 8.3 The 3:2:1 crack spread for Brent crude between June 2012 and May 2013 [7]
8.2 Process Development Driven by Market Demand 137
Summer Winter
(1,2)
U.S. Refinery Yields, vol% Jul 2014 Dec 2014
Liquified Petroleum Gases (LPG) (3) 5.3 2.4
(4)
Motor Gasoline 44 46.5
Aviation Gasoline 0.1 0.1
Kerosene-Type Jet Fuel 9.7 9.7
Kerosene 0.1 0.2
Distillate Fuel Oil 29.1 30.7
Residual Fuel Oil 2.4 2.3
Petrochemical Feed (naphtha) 1.3 1.4
Petrochemical Feed (other) 0.7 0.5
Special Naphthas 0.3 0.2
Lubricants 1 0.9
Waxes 0.1 0
Petroleum Coke 5.5 5.3
Asphalt and Road Oil 2.3 1.7
Still Gas (3) 4.2 4
Miscellaneous Products 0.6 0.5
Volume Gain (5) 6.7 6.4
(1) After removal of sulfur, nitrogen, and other contaminants.
(2) Volumes of solids are expressed in fuel-oil equivalents.
(3) Still gas is largely methane but can also contain C2-C4 compounds
(4) Does not include ethanol.
(5) The volume swells due to cracking. Light products have lower
densities than crude oil.
Fig. 8.4 Industry-average volume-percent yields from petroleum refineries in the United States.
Sulfur and other contaminants are excluded. Motor gasoline yields do not include ethanol. In
Europe, yields of motor gasoline and distillate fuel oil would be approximately reversed, and yields
of petroleum coke would be lower. About 90 vol.% of the products are fuels—gasoline, kerosene,
jet and fuel oils. Most refinery gases and petroleum coke also are used as fuels
• Assume the NKG gap is 23.9%—the same as for Brent in Table 8.3.
• Assume the crack spread is $11 per barrel.
• On this basis, the U.S. conversion incentive is $47.3 million dollars per day.
Today’s modern refineries waste almost nothing. It is possible to convert more
than 99 wt% of a conventional crude into something other than ash, and the ash can
be blended into clinker for cement manufacturing.
Figure 8.4 presents a quantitative petroleum product breakdown in the United
States for July and December 2014 [8]. About 83 wt% (90 vol.%) went to fuels. In
Europe and Asia, yields of motor gasoline and distillate fuel oil (diesel) are approx-
imately reversed.
138 8 Petroleum Processing and Refineries
According to the U.S. Energy Information Administration (EIA), the top ten refineries
in the U.S. have the capacities are listed in Table 8.5. Many of the largest refineries
are no longer owned by large integrated oil companies. In fact, many large plants
previously owned by big oil companies have been sold to smaller refining companies.
The location of the refineries stays the same as refineries cannot be easily relocated.
8.3 Refinery Capacities in the United States … 139
Table 8.5 Top 10 U.S. refineriesa operable capacity (EIA. Jan. 2018 database) [9]
Name of refinery Location Barrel per calendar day (BPCD)
Port Arthur Refinery (Motiva Port Arthur, Texas 636,500
Enterprises)
Galveston Bay Refinery Texas City, Texas 571,000
(Marathon Petroleum)
Baytown Refinery Baytown, Texas 560,500
(ExxonMobil)
Garyville Refinery (Marathon Garyville, Louisiana 556,000
Petroleum)
Baton Rouge Refinery Baton Rouge, Louisiana 502,500
(ExxonMobil)
Whiting Refinery (BP) Whiting, Indiana 430,000
Lake Charles Refinery (Citgo) Lake Charles, Louisiana 427,800
Beaumont Refinery Beaumont, Texas 344,600
(ExxonMobil)
Port Arthur Refinery (Valero) Port Arthur, Texas 335,000
Pascagoula (Chevron) Pascagoula, Mississippi 330,000
a Only
refineries with atmospheric crude oil distillation capacity
BPCD—barrels per calendar day
Table 8.6 lists the ten U.S. states with the largest refining capacity, comparing
data for 2005 with 2018. Note how capacities decreased during this time frame in
California, Pennsylvania, and New Jersey while they increased in the other seven. In
contrast, Wyoming has 4 refineries, their total capacity is only 105,000 barrels per
calendar day.
The majority of world’s 10 largest refineries are located in Asia, led by the 1.24
million bbl/day capacity Jamnagar refinery complex in India, followed by one each
in Venezuela, United Arab Emeritus (UAE) and Singapore, and three each in South
Korea and the United States. Table 8.7 lists the top 10 large oil refineries in the world,
based on processing capacity.
Many companies own several refineries, large and small. Based on 2015 data,
the five largest oil refining companies are: ExxonMobil Refining and Supply
Co. (5,580,000 BPCD), Royal Dutch Shell Group (4,109,000 BPCD), Sinopec
(3,971,000 BPCD), BP PLC (2,859,000 BPCD) and Saudi Arabian Oil Co. (2855
BPCD), where BPCD stands for barrel per calendar day.
Table 8.8 compares the refining capacity in thousand barrels per calendar day
(MBPCD) of the top 13 countries in the world in 2010 and 2014. Some countries do
not have any refineries. On the other hand, it can be seen that the capacity does not
necessarily reflect the demand inside the country. The refined products are imported
by the countries that need them. The United States, Canada, and certain other coun-
tries import products at some locations and export products from others.
140 8 Petroleum Processing and Refineries
Table 8.6 Ten U.S. states with the largest refining capacity [9]
State Number of operating refineries Crude distillation capacity (1000
BPCD)
2005 2018 2005 2018
Texas 27 29 4628 5701
Louisiana 17 17 2773 3303
California 21 16 2027 1892
Illinois 4 4 896 999
Washington 5 5 616 638
Pennsylvania 5 4 770 601
Ohio 4 4 551 598
Oklahoma 5 5 485 522
Indiana 2 2 433 442
New Jersey 5 3 666 418
Top 10 total 93 91 13,845 15,114
U.S. total 144 140 17,125 18,598
BPCD—barrels per calendar day
Table 8.7 Ten largest refineries in the world (2018) (in MBPCD: thousand barrels per calendar
day) [10]
Name of refinery Location Country Crude distillation capacity
(MBPCD)
Jamnagar Refinery Jamnagar, India 1240
(Reliance) Gujarat
Paraguana Refinery Punto Fijo, Venezuela 940
complex (PDVSA) Falcon
SK Energy Ulsan Refinery Ulsan South Korea 840
(SK energy)
Ruwais Refinery (Abu Ruwais UAE 817
Dhuba Natinal Oil
Company)
Yeosu Refinery (GS Yoesu South Korea 730
Caltex)
Ornan Refinery (S-oil) Onsan South Korea 669
Jurong Island Refinery Jurong Island Singapore 605
(ExxonMobil)
Port Arthur Refinery Port Arthur, United States 603
(Saudi Aramco) Texas
Galveston Bay Refinery Texas City, United States 571
(Marathon petroleum) Texas
Baytown Refinery Baytown, Texas United States 560
(ExxonMobil)
8.3 Refinery Capacities in the United States … 141
Table 8.8 Refining capacity by country (in MBPCD: thousand barrels per calendar day) [11]
Country 2010 2014
United States 17,736 17,791
China 10,302 14,098
Russia 5511 6338
India 3703 4319
Japan 4291 3749
South Korea 2712 2887
Saudi Arabia 2107 2822
Brazil 2093 2235
Germany 2091 2060
Iran 186p 1985
Italy 2396 1984
Canada 1951 1965
Mexico 1702 1375
United Kingdom 1757 1368
Although China was among the first countries to discover and use oil and natural
gas, the development of the modern Chinese oil industry was inhibited by relatively
low reserves and lack of demand from the transportation sector in earlier years.
Between 1966 and 1976, when petroleum use in Western countries was growing
exponentially, the Cultural Revolution quashed Chinese economic growth and, con-
sequently, the growth of energy demand. China began to drill wells in 1878, but
until 1949 only a few small oil fields were exploited. Since 1949, several large fields
have been discovered and exploited. Refining capacity also increased steadily. Today,
China has the second largest refining capacity in the world. Many Chinese refineries
are now as advanced as any in the world.
The United States and Europe contain 45% of the world’s refineries and account
for 46% of world’s crude distillation capacity. The high usage of FCC in North and
South America reflects the higher demand for gasoline in those regions. Note the
low use of thermal cracking and the high use of hydrotreating in the United States
and the Asia Dragons (South Korea, Singapore, Hong-Kong and Taiwan).
The goals of modern refinery operations are to be safe, economically feasible, and
environmentally acceptable. They can be divided into the following categories:
• Separation processes involve no change in the size and basic structure of
molecules. In other words, they are only physical separations, no chemistry is
142 8 Petroleum Processing and Refineries
For refiners, the goal of a conversion process is to reduce boiling point, primarily
by transforming low-valued heavy fractions into high-valued light products. This
is accomplished by increasing hydrogen-to-carbon (H/C) ratio through hydrogen
addition or carbon (coke) rejection. The most common hydrogen addition pro-
cesses are hydrotreating and hydrocracking, which increase the H/C ratio by adding
hydrogen to the molecules. Taken together, these two processes are referred to as
hydroprocessing. When discussing hydroprocessing, some authors include catalytic
reformers (which generates hydrogen) and isomerization units, in which hydrogen is
not consumed but is needed to reduce catalyst coking. The flow schemes for fixed-bed
hydrotreating and hydrocracking units are similar. Hydrotreating is an essential pre-
8.5 Conversion Strategy 143
slurry-phase hydrocrackers can. Products from the latter include naphtha, middle distillates, and
FCC feed
cursor to fixed-bed gasoil hydrocracking. Feeds and products for typical hydrotreaters
and hydrocrackers are shown in Table 8.9.
Other conversion processes—FCC, thermal cracking, and delayed coking – often
are called coke rejection processes. In fact, they yield both lighter molecules with
higher H/C and heavier products with lower H/C ratios at the same time. In coking and
FCC, the heavier product is coke, in which the H/C ratio is <1.0. It is common to hear
“carbon rejection” instead of coke rejection, but “carbon rejection” is a misleading
phrase; elemental carbon per se is not removed.
Refiners define conversion as any means of transforming material that boils above
a particular cutpoint, known as the conversion cutpoint, into material that boils below
the conversion cutpoint. A typical conversion cutpoint for naphtha-oriented opera-
tions is 400 °F (204 °C), while a typical conversion cutpoint for diesel-oriented
operation is 650 °F (343 °C). The equation for conversion is simple:
Table 8.10 Molecular weight, H/C and boiling points for selected hydrocarbons
Molecular Boiling point
Compound Weight Formula H/C °C °F
Paraffins
Methane 16.04 CH4 4.0 −164 −263.2
Ethane 30.07 C2 H6 3.0 −88.6 −127.5
Propane 44.10 C3 H8 2.67 −42.1 −43.7
Butane (iso) 58.12 C4 H10 2.50 −6.9 19.6
Octane (iso) 114.23 C8 H18 2.25 99.2 210.6
Cetane (n) 226.44 C16 H34 2.13 287 548.6
Aromatics
Benzene 78.11 C6 H6 1.0 80.1 176.2
Naphthalene 128.17 C10 H8 0.8 218 424.4
Benzopyrene 252.32 C20 H12 0.6 – –
where P = the amount of material in the product that boils below the conversion
cutpoint and FF = the total amount of fresh feed. In “true conversion” calculations,
the amount of P in the FF prior to conversion is subtracted from the denominator.
Table 8.10 compares compounds with different hydrogen content. For any given
class of hydrocarbons, “light” means a lower molecular weight, lower boiling point,
lower density, and a higher hydrogen-to-carbon ratio (H/C). Methane, the lightest
hydrocarbon, has an H/C ratio of 4.0. Benzopyrene has an H/C ratio of 0.6. The H/C
ratio of common crude oils ranges from 1.5 to 2.0, while the H/C ratio for asphaltenes
is about 1.15.
version and integrated. Almost all U.S. and European refineries are either conversion
or integrated refineries, as are the newer refineries in Asia, the Middle East, South
America, and other areas experiencing rapid growth in demand for light products. By
contrast, much of the refining capacity in Europe and Japan is in hydroskimming and
conversion refineries [12]. Due to shutdowns of many isolated refineries in Japan,
refining in that country is largely integrated with chemical plants within major indus-
trial zones near Chiba, Yokohama, and Osaka/Mizushima. The same could be said
for Europe, where refineries in Rotterdam/Europoort, La Mede/Laverra, Austria and
Italy are closely associated with chemical plants.
The simplest refinery configuration is topping refinery, shown in Fig. 8.5, which is
designed for petrochemical manufacture or production of industrial fuels for local
markets in remote oil-producing areas to reduce transportation costs. Only the desired
products are shown on the right side of the figure. Gas oil product can be heating oil,
fuel oil and lubricant feedstock.
146 8 Petroleum Processing and Refineries
A topping refinery consists of tankage, a distillation unit, recovery units for gas and
light hydrocarbons and necessary utility systems (steam, power and water treatment
plant). The refinery simply separates crude oil into light gas and refinery fuel gas,
naphtha (gasoline boiling range), distillates kerosene, jet fuel, diesel and heating oils,
and residual or heavy fuel oil. There is no hydrotreating unit to control sulfur levels
in the fuel products. Hence, topping refineries are suitable for light crudes or streams
that contain small amounts of sulfur components. The residual sulfur components in
products can be treated by solvent sweetening for removal.
In modern times, except for small refineries which make asphalt or in extremely
isolated locales in 3rd world countries, topping refineries are obsolete due to product
sulfur requirements.
With the addition of hydrotreating, reforming and product blending units to a top-
ping refinery, shown in Fig. 8.6, it will become a hydroskimming refinery which can
produce desulfurized distillate fuels and high-octane gasoline for market needs. In
this design, over half of the output can be residual fuel oil due to lack of conver-
sion. N-butane in LPG can be added into gasoline to increase RVP for winter use in
cars (shown as a dotted arrow as an optional flow). A typical hydroskimming refin-
ery would include an atmospheric distillation tower (pipestill), a catalytic naphtha
reforming unit, light-ends recovery and fractionation, treating, and blending. The
pipestill performs initial distillation of crude oil into gas, naphtha, distillates and
residual fuel. The naphtha may be separated into gasoline blending stocks, solvents
and reformer feed. The distillates include kerosene, jet fuel, diesel and heating oil.
The residual fuel is blended for use as a bunker fuel oil. Hydroskimming refineries
are commonplace in regions with low need for conversion, either due to light feed-
stocks (crudes) or market demand. In other words, there is no need to alter the natural
yield patterns of the crudes they process.
A conversion refinery has the most complex, versatile and efficient configuration.
A conversion refinery incorporates all of the basic units of topping and hydoskim-
ming refineries, plus gas oil conversion plants (catalytic cracking and hydrocrack-
ing), olefin conversion plants (alkylation and polymerization) and residue conversion
units (coking, RFCC, etc.). Many such refineries also incorporate solvent extraction
processes for manufacturing lubricants and petrochemical units to recover ethylene,
propylene, benzene, toluene and xylenes (BTX) for further processing into solvents
and high polymers.
8.6 Refinery Categories 147
Conversion refineries produce a whole range of gas, fuel and lubricant base oil
products. They improve the natural yield patterns of the crudes they process as needed
to meet market demands for light products, but they still produce some heavy, low-
value products, such as residual fuel and asphalt.
In the recent past, one could say that there were two types of conversion refineries:
the Cracking Refinery, shown in Fig. 8.7 for catalytic cracking, and the Coking Refin-
ery, shown in Fig. 8.8. Cracking refineries transform heavy crude oils to gasoline, jet
fuel, diesel fuel, and petrochemical feedstocks through catalytic cracking or hydro-
cracking or both. Coking refineries include coking in addition to catalytic cracking
and hydrocracking to convert gas oil and residuum fractions. Coking units “destroy”
the heaviest and least valuable crude oil fraction (residual oil) by producing cokes
and lighter streams that serve as additional feed to other conversion processes (e.g.,
catalytic cracking) and to upgrading processes (e.g., catalytic reforming) that produce
the more valuable light products. In general, all product streams are hydrotreated to
remove residual sulfur and nitrogen (hydrotreaters are not shown in the figure).
Refinery size is usually measured in terms of atmospheric distillation capacity,
as discussed in Sect. 8.3. In 1960, Wilbur L. Nelson developed Complexity Index
relative to the atmospheric (initial) distillation to quantify the relative costs of the
148 8 Petroleum Processing and Refineries
components that constitute the refinery (refinery units), listed in Table 8.11. The index
is used not only as an indicator of the investment cost but also the value addition
potential of a refinery. The sum of the products of the Nelson complexity index with
the capacity of each unit (or the percentage of crude oil it processes) in the refinery
is Equivalent Distillation Capacity (EDC) of the refinery in total as the benchmark
of investment cost and manpower requirement.
The process flow schemes for a representative high-conversion refinery are shown in
Figs. 8.11 and 8.12. We say “representative” because no two commercial refineries
are exactly the same. They contain different types of process units, and the units are
arranged differently. The configuration of a high-conversion refinery is very complex.
Many units are connected to each other, as some product streams of some units are the
feeds for the subsequent units. The processes for distillation cuts other than residua
(atmospheric and vacuum) are shown in Fig. 8.11, and the processes for residua are
shown in Fig. 8.12.
Prior to being subjected to refinery operations, crude oils are stored in storage
tanks (tank farm) within or adjacent to the refinery. Different grades of crude oils are
usually blended to meet the refinery feed requirement or specifications and refinery
constraints. For example, high acid crudes are blended with low acid crudes according
to the corrosion tolerance of the refinery equipment.
8.7 Refinery Configuration 151
The first major process unit in a refinery is the desalter, which removes dirt,
water, salt, and other water-soluble contaminants from blended crude oils. From
the desalter, the crude(s) go to an atmospheric distillation unit (ADU), which
produces several streams. These include refinery fuel gas containing mostly methane
and ethane, petroleum gas (C3 and C4 ) containing mostly propane and butanes,
one or two straight-run naphtha streams, one or two straight-run middle distillates
streams, and atmospheric residue (AR). All ADU products contain sulfur, either
as H2 S, inorganic sulfur, or organic sulfur compounds. The heavier fractions also
contain nitrogen, oxygen, and trace contaminants. At some point, with one major
exception, hydrotreating removes organic sulfur and other contaminants from C5 +
liquid streams. Sulfur can be removed from relatively sweet naphtha with mercaptan
oxidation (Merox) technology.
Here are the main destinations for individual streams from atmospheric distillation
units, from the bottom of the tower to the top:
• The AR goes to a vacuum distillation unit (VDU), which produces vacuum gas
oil (VGO) and vacuum residue (VR). Refiners allocate VGO based on business
requirements. Common VGO destinations include a fluid catalytic cracker (FCC),
a hydrocracker, a lube base stock facility, and fuel-oil blending. VR can be a
coker feed.
• Either before or after hydrotreating, straight-run middle distillates are divided into
light and heavy fractions, which become kerosene or gas oil, respectively. Both
152 8 Petroleum Processing and Refineries
C4= 2
Offgases C3=
11
12 13 Catpoly
1
i-C4
Multiple Offgas 2
C3=
Processing Plants
H2 from
SMR
H2S Alkylation
10
C5 2
Amine/Claus
C6 1
Naphtha 2 Isomerization
Splitter
2
3 C6 to C10
14
Catalytic
Atmospheric Reforming
Naphtha HT
Distillation 3
2
6 7 8
4
FCC 2 3
Intermediate 5 6
6
Gas Oils
VGO HT 9
7
6
AR VR Vacuum
Distillation
From Resid Processing
DAO, CGO, VBGO
To Residue Upgrading
- Visbreaking
1 – Hydrogen 88 – Olefin Plant Feed
- Coking
- Solvent deasphalting 2 – Gasoline 9 – Fuel Oil, Resid Upgrading
- Resid FCC 3 – Heavy Naphtha 10 – Sulfur
10
- Resid hydroprocessing – Kerosene / Jet
4 11 – Refinery Fuel Gas
11
5 – Gas Oil / Diesel 12 – Propane
12
HT = Hydrotreating – Intermediate Gas Oil 13
6 13 – Light Olefins
= Offgases
7 – Lube Base Stock 14 – Aromatics
14
Fig. 8.11 Oil refinery process units: upgrading processes for top-of-the-barrel cuts—VGO and
lighter—which boiling below about 1050 °F (556 °C). Please refer to the key for stream descriptions.
Almost all units produce some kind of offgas; offgas processing is described above. No single
drawing can show all the complexities of a high-conversion refinery. Polymerization unit for poly
gasoline and a visbreaker are not shown in this figure. Please refer to the text for additional details
8.7 Refinery Configuration 153
HVGO
AR DAO
VR
C3=, C4=
AR
Gasoline Solvent
Resid HT Cycle oil Deasphalting
with RFCC Slurry oil
Asphalt
Naphtha
AR Visbreaker gas oil AR Mid distillates
VR UCO
Ebullated Bed
Visbreaking Hydrocracking
Sediment
Fuel Oil
Coking Slurry
Delayed, Fluid Hydrocracking Sediment
Coke Pitch
Fig. 8.12 Oil refinery conversion processes for bottom-of-the-barrel cuts, which can include either
atmospheric residue (AR) and/or vacuum residue (VR). Commercially, RFCC processes mostly
AR Ebullated-bed hydrocracking processes either AR or AR/VR blends. Slurry hydroprocessing
handles VR, coal, FCC slurry oil, and/or coal tar. Delayed coking processes VR. Please refer to the
text for additional details
can serve as fuel oil. If it meets specifications, the kerosene can be blended into
jet fuel and the gas oil can be blended into diesel. Certain refineries produce just
one middle distillate stream, typically wide-range gas oil.
• Naphtha is split into light and heavy fractions. Desulfurized light naphtha (LN)
includes C5 molecules along with differing amounts of C4 and C6 . LN can be
blended into gasoline or separated into constituent components. Small amounts of
mixed butanes can be blended into gasoline. N-butane is sold as such or isomerized
into i-butane. I-butane goes to alkylation (ALKY) units or is dehydrogenated to
C4 olefins for chemical plants. Without excess isobutane, C3 –C5 olefins can go to
a catalytic polymerization unit to produce gasoline. Heavy naphtha (HN) includes
C6 –C12 molecules. Catalytic reforming (CRU) and isomerization (ISOM) units
are common HN destinations. C6 compounds may be excluded from CRU feeds if
a refiner wants to minimize benzene production. C9 –C12 molecules can be blended
into aviation gasoline.
• H2 S is removed from refinery fuel gas by adsorption in alkanolamine units. The
H2 S is converted to elemental sulfur in Claus process units. Sweet refinery fuel
154 8 Petroleum Processing and Refineries
gas serves as fuel gas throughout the refinery. Petroleum gas (C3 and C4 ) is sold as
liquefied (LPG) or exported to a chemical plant, either before or after segregation
into individual components.
• Thermal cracking units, such as cokers and slurry-phase hydrocrackers, produce
significant quantities of methane and ethane. In catalyst-based conversion unit-
s—FCCs and fixed-bed hydrocrackers—C1 /C2 production is low, and for paraf-
fins, the iso/normal ratio is high. Processes which operate without external hydro-
gen—FCC and coking—generate olefins, aromatics, hydrogen and coke. FCC is
a primary source of propylene and butylenes for downstream chemical plants.
In hydrocracking units, which operate with large amounts of external hydrogen,
olefins and coke are low, product aromatics are lower than feed aromatics and light
alkanes including isobutane for alkylation feed. The polyolefin feeds, ethylene,
propylenes, butylenes and butadiene, are mainly produced from steam cracking.
• Visbreaking achieves some conversion, but its primary purpose is to reduce the
viscosity of heating oils and fuel oils, making them suitable for furnaces and
improving their flow in tubes and pipes.
As much as anything else, VR conversion determines refinery profitability.
Options are shown in Fig. 8.12. The cost of crudes depends on their sulfur and
VR content. High-sulfur, high-VR crudes cost less, so refineries configured to han-
dle them tend to be more profitable. VR is a source of asphalt, bright-stock lubri-
cants, greases, waxes, and specialty carbon products. To a certain extent, VR can be
blended into No. 6 fuel oil (bunker fuel). Processes for converting AR and VR into
lighter products include delayed coking, visbreaking, modified versions residue
FCC (RFCC), and residue hydrocracking with either ebullated bed or slurry-phase
technology.
Individual process units can be significantly different, but they have several things
in common:
• Heat. All units require heat, which is supplied by fuel gas or fuel oil, but sometimes
just by steam. Even if external heat isn’t necessary during normal operation of a
particular unit, heat is always required for startup. In high-conversion refineries,
off-gases from operating units might provide all of a refinery’s fuel gas needs.
Others refineries are “fuel gas long,” meaning that they generate more gas than
they can legally burn in a flare.
• Power and Steam. All units require electric power and most use one or more
grades of steam. FCC units and sulfur plants are net producers of high-grade
steam, but they usually import low-grade steam from elsewhere in the refinery.
• Amine Adsorption: Recovery of Sulfur and CO2 . H2 S-containing “sour” off-
gases, both olefinic and non-olefinic, are treated in multiple plantwide amine sys-
tems, which remove the H2 S (and any CO2 which happens to be present) by
acid/base adsorption and transport them the refinery sulfur plant. Sulfur plants
employ the Claus process, in which H2 S is converted to elemental sulfur via com-
bustion. Claus tail-gas plants boost recovery to >99.9%.
• Offgas Processing. Offgas is a general term for gases produced by refinery units.
Depending on their source, offgases can include H2 , N2 , H2 S, light paraffins
8.7 Refinery Configuration 155
(methane, ethane, propane and butanes), and light olefins (ethylene, propylene,
and butenes.) As mentioned, H2 S is removed from all gases by amine adsorption
and converted into elemental sulfur. The purified “sweet” gases go to either to a
saturated gas plant (sat gas plant) or an unsaturated gas plant (unsat gas plant).
The sat gas plant treats offgases from hydrotreaters and catalytic hydrocrackers,
and it can also co-process natural gas. Feeds to the unsat gas plant contain ethy-
lene, propylene, and butylenes; most olefins are produced by FCC and coking
units. Products from the gas plants become refinery fuel gas, which contains H2 ,
N2 , ethane, and/or ethylene; LPG—liquid petroleum gas, which contains mostly
propane with limited amounts of butanes; and one or more refinery olefin streams.
Refinery fuel gas is burned to supply process heat. LPG can be sold as-is or frac-
tionated to purify the propane. Isobutane goes to refinery alkylation units, where
it reacts with propylene and butylenes to form high-value C7 –C8 alkylate, excel-
lent gasoline blend stock. Purified ethylene and propylene go to polyolefin plants.
Refinery propylene is a primary source of polypropylene.
• Hydrogen is required by hydrotreaters, hydrocrackers, isomerization units, cat-
alytic dewaxing units, and others. Hydrogen is produced by catalytic reformers,
steam/methane reformers, and nearby olefin plants. FCC and delayed coking units
also produce significant amounts of low-purity hydrogen, but it is costly to recover
and use that hydrogen.
• Over-the-fence gases. It is common in certain areas for refineries to purchase
“over-the-fence” hydrogen and nitrogen from industrial gas suppliers. Nitrogen is
required during shutdowns and startups as an inert gas for protection of the interior
of the equipment.
• Instrument Air. Valves are opened and closed with actuators driven by “instru-
ment air.” It is important to keep the air clean and dry to prevent the fouling of
lines.
Figure 8.13 illustrates how streams from individual process units might be blended
into four major groups of finished fuel products. (A fifth major group—lubricants—is
discussed separately.) As shown, the gasoline blender prepares finished gasoline
from 12 different streams. Oxygenates (ethanol) are added separately to the blender
product, called Reformulated Gasoline Blendstock for Oxygen Blending (RBOB
gasoline), commonly known as reformulated gasoline (RFG). In some regions, such
as Southern California, detergent additives are added at blending and distribution
terminals, some of which are located hundreds of miles away from any refinery. The
figure over-simplifies middle distillate blending.
Kerosene and jet fuel are shown in the same box, because their boiling ranges are
similar. The cut points of certain streams are changed based on market demand, and
some streams can go to two or three places. Depending on how it is cut, straight-run
light gas oil can become jet fuel, kerosene, or diesel.
156 8 Petroleum Processing and Refineries
Refineries can produce several grades of fuel oil. The bottom line is: A material
isn’t gasoline unless it meets a specification, such as ASTM D-4814 in the United
States and many other countries, or EN 228 in Europe.
References
1. Berthelot M Collection des anciens alchimistes grecs (3 vol, Paris, 1887–1888, p 161); F.
Sherwood Taylor, “The Origins of Greek Alchemy,” Ambix 1 (1937), 40
2. Energy Timelines. Energy Kids. http://www.eia.gov/petroleum/. Retrieved 3 Dec 2015
3. Pees ST (2004) Whale oil versus the others. In: Oil history, Samuel T. Pees, Meadville, PA
4. Frank F (2007) Oil Empire: visions of prosperity in Austrian Galicia. Harvard University Press
5. Hsu CS, Robinson PR (2006) Practical advances in petroleum processing. Springer, New York,
NY
6. OSHA Technical Manual (2005) Section IV, Chapter 2, “petroleum refining processes”. U.S.
Department of Labor, Occupational Safety and Health Administration, Washington, DC
7. Falling crack spreads cloud near term for refiners. http://seekingalpha.com/article/1468771-
falling-crack-spreads-cloud-near-term-for-refiners. Retrieved 3 Dec 2015
8. U.S. Energy Information Administration: Detailed breakdown. http://www.eia.gov/dnav/pet/
pet_pnp_pct_dc_nus_pct_m.htm. Retrieved 3 Dec 2015
9. (a) U. S. Energy Information Administration: Top 10 US Refineries Operable Capacity. Jan.
1, 2018 database, published June 2018. http://www.eia.gov/petroleum/refinerycapacity, (b)
https://www.eia.gov/petroleum/refinerycapacity/table4.pdf. Retrieved 30 Sept 2018
10. List of Oil Refineries. https://en.wikipedia.org/wiki/List_of_oil_refineries. Retrieved 1 Oct
2018
11. (a) Asia-Pacific refining primed for capacity growth. Gas Oil J 112(12), Dec. 1, 2014. pp
34–45; (b) https://www.statista.com/statistics/273579/countries-with-the-largest-oil-refinery-
capacity/. Retrieved 1 Oct 2018
12. An Introduction to Petroleum Refining and the Production of Ultra Low Sulfur Gasoline and
Diesel Fuel. The International Council on Clean Transportation, 24 Oct 2011
13. Hauge K (2015) Refining ABC. http://www.statoil.com/en/InvestorCentre/Presentations/
Downloads/Refining.pdf. Retrieved 3 Dec 2015
Chapter 9
Crude Storage, Blending, Desalting,
Distillation and Treating
Large quantities of crude oils are transported through pipelines or tankers. At the
marine terminal, a cargo of crude oil may be routed through a pipeline directly to a
storage tank in the refinery tank farm, or transferred to holding tank, where it is kept
temporarily before going to the refinery. The marine terminal also has berths to load
refined products. Storage tanks are containers that hold the crude oil or compressed
gases (gas tanks), as well as intermediate stocks (partially refined), finished products
and chemicals for the short- or long-term storage.
Crude oils shipped to the refinery are stored in storage tanks. Liquid storage tanks
are often cylindrical in shape, perpendicular to the ground with flat bottoms, with
a fixed frangible or floating roof. Gas tanks for compressed or liquefied gas are in
spherical shape.
Different grades of crude oils are stored in different tanks. Care must be taken
to avoid mixing incompatible crudes, which can cause precipitation (due to deas-
phalting)., and fouling. Above ground storage tanks can also be used to hold blended
crudes, refined products, water, waste matter, and hazardous materials, while meeting
strict industry standards and regulations.
A refinery seldom refines a single crude. Prior to being charged to the refinery
units, crude oils and imported stocks are unloaded to storage tanks, then blended and
transferred to surge tanks for individual units. Proper blending improves distillation
unit throughput and the performance of downstream units. It also can improve product
quality and reduce energy cost [1]. Crude blending is based on computer models,
which are used in conjunction with scheduling models and operations plans. The
models and plans include timing, desired volumes, etc. They are unique for every
refinery, because they depend on the refinery configuration, logistics constraints,
tank inventory, feedstock composition, and local (and global) market forces. For
operations planning, some refiners still use comparatively simple stand-alone LP
© Springer Nature Switzerland AG 2019 159
C. S. Hsu and P. R. Robinson, Petroleum Science and Technology,
https://doi.org/10.1007/978-3-030-16275-7_9
160 9 Crude Storage, Blending, Desalting, Distillation and Treating
9.2 Desalting
If the crude isn’t desalted, residual solids can clog downstream equipment and
deposit on heat exchanger surfaces, thereby reducing heat-transfer efficiency. Salts
can induce corrosion in major equipment and deactivate catalysts.
9.3 Distillation
Distillation can be considered as the heart of any refinery. Crude oils are made
of numerous components of different boiling points. The simplest way to separate
them is with continuous distillation into different fractions (distillates or cuts) of
various boiling ranges. At just above atmospheric pressure, heavy molecules in most
crude oils decompose above 650 °F (~350 °C). To achieve additional separation of
heavy fractions, continuous distillation is carried out under vacuum (i.e., reduced
pressure-typically at 40 mmHg) to boil out additional components. Vacuum distil-
lation increases the yield of total distillates. The relationship between boing points
under atmospheric pressure and under 40 mmHg vacuum is shown in Fig. 2.1. For
9.3 Distillation 163
example, at 500 °F under 40 mmHg vacuum, the compounds with boiling point
at 750 °F under atmospheric pressure can be boiled out. Hence, the atmospheric
equivalent boiling point (AEBP) at 500 °F under 40 mmHg vacuum is 750 °F. The
AEBP of 650 °F is ~900 °F. Hence, additional components that boil between 650
and 900 °F can be distilled out of the crude oil under 40 mmHg vacuum without
severe decomposition. The upper temperature for vacuum distillation in refineries
can be slightly higher, up to 1050 °F. At this temperature, there is a greater tendency
for thermal decomposition, but the decomposition does not occur immediately; the
feed flows out of the column and undergoes cooling before any damage is done.
Figure 9.4 shows a simplified diagram for crude oil distillation. There are many
trays in a distillation tower (also called a pipestill or column), which will be discussed
later. The desalted crude is introduced near the bottom of the atmospheric distillation
tower. The lightest fractions, gases and naphtha, flow out of the tower at the top as
an overhead stream, which can be further fractionated into separate gas and naphtha
streams. The effluents from middle trays are heavy naphtha, kerosene, light gas oil and
heavy gas oil. Kerosene and light gas oil are often referred also as middle distillates
or distillate fuels that include kerosene, jet fuel and diesel. Steam is introduced to
164 9 Crude Storage, Blending, Desalting, Distillation and Treating
LPG, Propane
Condenser
Reflux Drum
Naphtha
Naphtha
Reflux Splitter
Atmospheric
Distillation Unit
Vacuum
Atmospheric Distillation Unit
Reboiler Residue
Vacuum Residue
overlaps
Tail ends
between naphtha and kerosene. The leading edge of the kerosene distillation curve
overlaps with the tailing edge of naphtha. The overlap between cuts can be “sharp”
or “sloppy,” depending on several factors, especially oil flow rates, steam flow rate,
and heat balance. The designations are arbitrary, and often nothing can be done to
decrease overlap at maximum flow rate without making hardware changes. Overlaps
exist for all distillation cuts, in some commercial units, the overlaps are very large.
Such “sloppy cuts” are more common in units running far above (or far below) their
design feed rate.
Figure 9.6 demonstrates the application of boiling point overlaps. The desired
cutpoints for light naphtha, heavy naphtha, kerosene, and heavy diesel are 90 °F
(32 °C), 190 °F (88 °C), 300 °F (149 °C), and 525 °F (274 °C), respectively. Due to
operational constraints, the observed effective cutpoints are 99 °F (37 °C), 188 °F
(87 °C), 302 °F (150 °C), and 523 °F (273 °C), respectively. The overlap between
kerosene and heavy diesel is considerable. If we move the initial boiling point (IBP)
for the heavy diesel at 360 °F (182 °C) instead of 523 °F (274 °C), it would result
in considerable entrainment of valuable kerosene into the far-less-valuable bottom
product.
A petroleum refinery can adjust distillation yields to meet market demands, in
part, just by adjusting cut points. The swing cut between 150 and 205 °C can go into
any of the three products—naphtha (gasoline), kerosene (jet fuel) and gas oil (diesel
and heating oil), shown in Fig. 9.7, depending on the seasonal and market demands.
166 9 Crude Storage, Blending, Desalting, Distillation and Treating
Heavy Diesel
Cutpoint
523°F (273°C)
Temperature, ºF
Kerosene
Heavy Naphtha Cutpoint
Cutpoint 302°F (150°C)
188°F (87°C)
Light Naphtha
Cutpoint
99°F (37°C)
Fig. 9.7 Swing cut region for naphtha, kerosene and gas oil
Figure 9.8 shows a flow diagram of an atmospheric distillation unit. The crude oil
enters a desalter at 250 °F to remove salt and water, as described before. The desalted
oil goes through a network of pre-heat exchangers to a fired heater, which brings the
temperature up to 657–725 °F (347–385 °C). If the oil gets much hotter than this,
9.3 Distillation 167
it starts to crack, generating carbon. The carbon would deposit inside the pipes and
equipment through which the oil flows.
The hot crude enters the tower just above the bottom, as shown in Fig. 9.8. Steam is
added at the bottom to enhance separation; it does so largely by decreasing the vapor
pressure of hydrocarbons in the column. When it enters the tower, most of the oil
vaporizes. The steam flows upward with vaporized crude while the condensed liquid
flows downward as in a countercurrent fashion. The hottest trays are in the bottom
section with the coolest at the top section. Unvaporized oil drops to the bottom of
the tower, where it is drawn off.
Products are collected from the top, bottom and side of the column. Side-draw
products are taken from trays where the temperature corresponds to the cutpoints for
a desired product (naphtha, kerosene, light gas oil and heavy gas oil). Some of the
side-draws can be returned to the tower as a pump-around or pump-back stream to
control tower temperatures and improve separation efficiency.
Two side cut strippers for naphtha and gas oil are shown as examples. There can be
additional strippers for kerosene (jet fuel) and diesel (light gas oil). Also not shown
is the reboiler at the bottom of the tower; this will be discussed later. An atmospheric
distillation tower usually contains 30–50 fractionation trays, with 5–8 trays in each
section. Product strippers for cut streams also have 5–8 trays. The strippers remove
entrained light components from liquids. Stripper bottom streams can be drawn off
as products of a specific boiling range or returned to the distillation tower.
Inside the distillation tower (also called pipestill or column), the vapors rise
through the distillation trays, which contain perforations, bubble caps, downcom-
ers, and/or modifications thereof, shown in Fig. 9.9 (perforations on the trays not
168 9 Crude Storage, Blending, Desalting, Distillation and Treating
Fig. 9.9 Bubble cap and downcomer on a distillation tray. Intermediate products are removed
through side-draw trays
shown). Vapors and liquids flow counter-currently. Each tray permits vapors from
below to bubble up through the relatively cool condensed liquid on top of the tray.
This vapor/liquid contact knocks heavy material out of the vapor. Condensed liquid
flows down through a pipe (downcomer) to the hotter tray below, where the higher
temperature causes re-evaporation. A given molecule evaporates and condenses many
times before finally leaving the tower.
Figure 9.10 is another drawing for an atmospheric distillation tower. The section
above the feed tray is called enriching, or rectification, section and the section below
the feed trap is stripping section. Gas and naphtha are withdrawn from the top tray
as overhead. After condensation, a portion of the liquid is introduced back to the
top tray as reflux. Reflux also controls temperature in the enriching section. It also
controls entrainment of heavier components in the naphtha (the lightest distillate).
Reflux ratio is the amount of reflux liquid returning to distillation tower divided by
the amount of liquid withdrawn as product per unit time. With a higher reflux ratio,
fewer theoretical plates are required.
At the bottom of the tower, the liquid (atmospheric residue) passes through a
reboiler to recover light components from the heavy liquid. The reboiler helps control
temperatures in the stripping section.
The bottom stream from the main fractionator (atmospheric distillation tower) is
called atmospheric bottoms, atmospheric residue, reduced crude, topped crude, or
long resid.
9.3 Distillation 169
Fig. 9.10 Atmospheric distillation (Fractionation) tower with a reflux for overhead and a reboiler
for bottoms
Figure 9.11 shows a flow diagram of a vacuum distillation unit. The atmospheric
residue goes to a fired heater, where the typical outlet temperature is about 730–850 °F
(390–450 °C). From the heater, the atmospheric residue goes to a vacuum distillation
tower. Steam ejectors reduce the absolute pressure to 25–50 mmHg vacuum, or about
7.0 psia (0.5 bara). Under vacuum, hydrocarbons vaporize at lower temperatures than
atmospheric boiling points. For example, the equivalent atmospheric boiling point of
800 °F under 40 mmHg vacuum is ~1050 °F. Thus, molecules with normal boiling
points above 650 °F (343 °C) are less likely to undergo thermal cracking and can
be vaporized at lower temperatures. There are fewer trays than the atmospheric
distillation tower to fractionate the topped crude into light vacuum gas oil, heavy
vacuum gas oil and vacuum residuum at the bottom. As in atmospheric distillation,
some gas and light components entrained or decomposed during heating in furnace
are carried out at the top of the tower.
170 9 Crude Storage, Blending, Desalting, Distillation and Treating
The products from distillation prior to upgrading are called straight-run products. At
a given set of cutpoints, the yields of different fractions depend on the crude oil being
processed. Figure 9.12 shows TBP of a light and a heavy crude for kerosene yield
(cut points between 315 and 450 °F). The light crude yields more kerosene than the
heavy crude, and hence has a higher value.
Table 9.2 lists a few crude oils and their typical straight-run yields. Total naphtha
includes light, medium and heavy naphtha, and the middle distillates include kerosene
and atmospheric gas oil. Naphtha is used for making gasoline and aromatics, kerosene
for jet fuel and atmospheric gas oil for diesel. Table 9.3 shows that the demand
for transportation fuels exceeds the straight-run yields for the crudes in Table 9.2.
Obviously, crudes containing less heavy material—VGO and vacuum residue are
more valuable.
The higher-valued crude oils, such as Brent and Bonny Light, have higher API
gravity with higher naphtha and middle distillate yields. They tend to have less sulfur.
The oil having high sulfur content increases processing costs because the sulfur must
be removed. Hence, the oils, such as Green Canyon and Ratawi, have lower values.
Since sulfur is not removed during distillation, the straight-run distillation products
have to be treated for sulfur removal.
Products from the crude distillation unit, i.e., the straight-run distillates, go to other
process units, as shown in Table 9.4. The lightest cuts are gas and light naphtha. The
gas goes to a gas processing plant or is liquefied into liquefied petroleum gas (LPG).
9.3 Distillation 171
Fig. 9.12 Comparison of kerosene yields from a light and a heavy crude oil [2]
The light naphtha can be hydrotreated and sent to the motor gasoline blending pool.
Heavy naphtha is a feed for catalytic reforming units.
Kerosene can be used for lighting, heating, and for making jet fuel. In either case,
it must first undergo hydrotreating. Light gas oil can go to diesel fuel (distillate fuel
oil) blending.
Heavy gas oil can become fuel oil, diesel, lube base stock, or a light component
of feed for fluid catalytic cracking (FCC) or hydrocracking.
Vacuum gas oil (VGO) and vacuum resid (residuum or residue, VR) are low
valued. They are normally converted into higher-value products through various
upgrading processes, as in a conversion refinery. VGO can become fuel oil or lube
base stock, but its primary destinations are FCC and hydrocracking units, which
are discussed in subsequent chapters. Figure 8.12 gives more details of the possible
destinations of vacuum resid which is also known as “bottom of the barrel”.
CH3 3 H2 CH3
toluene methylcyclohexane
+ 4 H2
phenanthrene sym-dinaphthenobenzene
+ 2 H2
naphthenonaphthalene
+ H2
dihydrophenanthrene
+ 3 H2
dinaphthenocyclohexane
Fig. 9.14 Thermodynamic calculations illustrating the competition between the saturation (hydro-
genation) and the condensation (dehydrogenation) of polyaromatics. Data for the graphs were
generated by Aspen Plus for a six component system comprising naphthalene (C10 H8 ), tetralin
(C10 H12 ), decalin (C10 H18 ), o-xylene (C8 H10 ), chrysene (C18 H12 ) and hydrogen (not shown)
176 9 Crude Storage, Blending, Desalting, Distillation and Treating
A
peri-condensaƟon
C
cata-condensaƟon
Fig. 9.15 Zig-zag mechanism for the condensation of polyaromatics by sequential addition of
2-carbon and 4-carbon units. The isomers shown are a naphthalene, C10 H8 ; b phenanthrene,
C14 H10 ; c pyrene, C16 H10 ; d benzo[e]pyrene, C20 H12 ; e benzo[ghi]perylene, C22 H12 ; f coronene,
C24 H12 ; g dibenzo[b,pqr]perylene, C26 H14 ; h benzo(pqr)naphtho(8,1,2-bcd)perylene, C28 H14 ;
i naphtho[2’.8’,2.4]coronene, C30 H14 ; and j ovalene, C32 H14 . Note how the H/C ratio goes down
as condensation increases, from 0.8 for naphthalene to 0.4375 for ovalene
CH3 CH4 H 2 H2
Alkylkation
(olefin addition) Cyclization
H
Fig. 9.16 Mechanism for the addition of rings to an existing layer of coke
9.5 Treating/Sweetening 177
9.5 Treating/Sweetening
Gases and straight-run distillates need to be treated for removal of undesirable impu-
rities before they can be used as products.
Several treating processes entail seemingly simple acid-base reactions, such as
alkanolamine treating and caustic scrubbing. Alkanolamine treating removes acid
gases—H2 S and CO2 —from fuel gas and off-gas streams. In some hydrotreaters and
hydrocrackers, high-pressure amine units remove H2 S from the recycle gas. Caustic
scrubbers are used in several ways, including removing the last traces of H2 S from
the hydrogen used for processes in which the catalysts are highly-sulfur sensitive.
9.6 Hydrotreating
Figure 9.17 presents a process flow scheme for a one-reactor fixed-bed hydrotreater
with four catalyst beds. Reaction conditions depend on feed quality and process
objectives.
The feed is warmed with heat exchange, mixed with hydrogen-rich gas and passed
through a furnace, where it is heated to the desired reactor inlet temperature, typically
600–780 °F. The temperature depends on process objectives and catalyst activity.
The heated mixture flows down through reactors loaded with catalysts. In many
diesel hydrotreaters and almost all VGO hydrotreaters, the reactors have multiple
beds, separated by quench decks. As the hydrotreating reactions occur, they con-
9.6 Hydrotreating 179
Treat
Gas
Recycle Gas
Furnace Compressor
Makeup H2
Quench
Gas Makeup Gas
Recycle Compressor
Gas
TC
Purge
TC
Gas
Reactor
TC
Amine Unit
H2S
(ads)
CHPS
F/E Heat LP Flash
Exchange Gas
Cooler
Fig. 9.17 Representative hydrotreating unit with one reactor, four catalyst beds, and two flash
drums. Naphtha hydrotreaters have one bed in one reactor, and may employ hydrogen once-through
(no recycle). F/E exchanger is the feed/effluent heat exchanger. Temperature controllers (TC) control
temperature by manipulating the flow of quench gas. CHPS is the cold high-pressure separator.
CLPS is the cold low-pressure separator. Wash water is injected to remove ammonia as aqueous
ammonium bisulfide, which otherwise would precipitate in cold spots in or downstream from the
CHPS, blocking flow and inducing corrosion
sume hydrogen and generate heat. The heat is controlled by bringing relatively
cool recycle gas into the quench decks, where it mixes with reaction fluids from
the bed above. Makeup gas comes into replace consumed hydrogen. Gas flow can
be once-through in naphtha hydrotreaters, but in distillate and VGO hydrotreaters,
unconsumed hydrogen is recycled.
Fluids exiting the reactor are cooled with heat exchange before going to an arrange-
ment of separation towers, which include a stripper and, for high-pressure units, two
or more flash drums.
Hydrotreating produces both H2 S and NH3 . Under reaction conditions, these
remain in the gas phase. But at lower temperatures, they combine to form solid
ammonium bisulfide (NH4 SH). Ammonia also reacts with chlorides to form NH4 Cl;
chloride can come with makeup gas, feed, or wash water. These salts can deposit in air
coolers and heat exchangers, blocking flow and—even worse—inducing corrosion.
180 9 Crude Storage, Blending, Desalting, Distillation and Treating
Fortunately, they are water-soluble, so they can be controlled by injecting wash water
into the reactor effluent for removal.
The CHPS flash gas is recycled. An optional amine unit removes H2 S from the
recycled gas. Some of the recycle gas might be purged to control hydrogen purity.
Depending on the unit and the feed, oils from the separation towers can go various
places. Stripped naphtha, kerosene or diesel might go directly to downstream units
or product blenders. In VGO or residue hydrotreaters, liquids from the separation
section go to a fractionator. Other off-gases, along with sour high-pressure purge gas,
are treated with amine before going to other units or the refinery fuel-gas system.
Olefins are rare in straight-run feeds, but they are relatively abundant in cracked stocks
from coking or FCC units. Saturation of olefins occurs rapidly. If not controlled, it
can lead to polymerization and consequent plugging of catalyst beds. The best ways
to control olefins are (a) to design hydrotreaters with small low-temperature beds up
front and/or (b) to employ activity grading. In activity grading, catalysts are layered,
with low activity catalysts on top, followed by successively more-active catalysts.
2 H2
CH3CH2-SH CH3CH3 + H2S
ethyl mercaptan Direct ethane
2 H2
CH3-S-C4H9 CH4 + C4H10 + H2S
t-butylmethylsulfide Direct methane isobutane
3 H2
CH3-S-S-CH3 2 CH4 + H 2S
Dimethyldisulfide (DMDS) Direct methane
2 H2
+ H 2S
Direct
S
dibenzothiophene biphenyl
2 H2
+ H2S
S Direct
H 3C CH3 H3C CH3
4,6-Dimethyldibenzothiophene
Indirect 2 H2
2 H2 + H2S
S
H 3C CH3 H3C CH3
after prior saturation, the crossover phenomenon affects the production of ULSD
significantly, so much so that it governs the design of commercial units.
Figure 9.19 shows representative HDN reactions, and Fig. 9.20 presents the mech-
anism for the HDN of quinoline. As with sulfur removal from hindered DMDBTs,
the aromatics crossover phenomenon is important for deep HDN because nitrogen
removal requires prior saturation of an aromatic ring adjacent to the nitrogen atom.
2 H2 C 3H 7
H2
N N NH2
fast slow
quinoline H
slow 2 H2 fast 3 H2 3 H2
3 H2 H2 C3H7
N N Hydrogenolysis NH2
H Rate limiting
Not reversible
H2
C 3H7
+ NH3
oxygen, a bio-derived oil in the same boiling range might contain >40% oxygen.
Transportation fuels must meet tight specifications, or they can’t be used—at least
not for very long. Producing fuels from bio-derived oils in conventional oil refineries
requires extra hydrogen and presents processing and storage challenges. Large-scale
hydrogen production generates large quantities of CO2 . When assessing the environ-
mental impact of replacing conventional crude oil fractions with bio-derived oils, it
is crucial to consider the hydrogen required for upgrading.
Metals and other trace elements poison catalysts in hydrotreaters and downstream
process units. The following are the most troublesome:
• Nickel and vanadium are present in high-boiling fractions, mostly in asphaltenes.
Asphaltenes are mixtures of waxy solids with porphyrins.
• Corrosion generates soluble iron.
• Entrained salt brings in alkali and alkaline earth salts, primarily carbonates and
bicarbonates of sodium and calcium.
• Arsenic is present in crudes from West Africa, the Ukraine, Canada, and elsewhere.
Synthetic crudes from oil sands and oil shale tend to contain significant amounts.
Arsenic is one of the worst catalyst poisons—hundreds of times worse, on a weight
basis, than Ni, V, or Fe. Arsenic forms organo arsines, which are relatively volatile.
They tend to be most abundant in middle distillates.
184 9 Crude Storage, Blending, Desalting, Distillation and Treating
The earliest hydrodesulfurization catalysts were bauxite and fuller’s earth. Later,
catalysts containing cobalt molybdate on alumina and nickel tungstate on alumina
substantially replaced the earlier catalyst and these catalysts are still used very exten-
sively.
Most commonly used hydrotreating catalysts are MoS2 promoted by Cox Sy and/or
NiS on gamma alumina. WS2 has also been used commercially in place of MoS2 .
Improved compositions and manufacturing methods have increased activity more
than ten-fold since the process was invented in the 1950s, but despite intense ongoing
efforts to find better catalysts, the original raw materials remain unsurpassed. Other
metal sulfides, such as RuS2 , IrS2 , OsS2 , and RhS2 , are more active than MoS2 and
WS2 , but they either deactivate too quickly or they are too expensive.
When the catalysts are manufactured, the metals are oxides, which have relatively
low activity. Before use, they must be activated by reductive sulfidation, a process
commonly known as sulfiding.
For the low-pressure HDS of light feeds, CoMo catalysts are preferred. For high-
pressure deep desulfurization and HDN, NiMo catalysts are used, either alone or in
combination with CoMo. Some suppliers claim that “sandwich” configurations, with
alternating layers of CoMo and NiMo catalysts, provide superior HDS performance.
In the past, catalysts based on Ni-promoted WS2 were employed for HDN, due to
the high saturation activity of the WS2 . However, WS2 catalysts are seldom if ever
used now.
The catalyst support is just as important as the active metals. The support must be
strong enough to endure considerable pressure and hydraulic stress, and it must have
a high surface area. The supports are microporous, with 99+ percent of their surface
9.6 Hydrotreating 185
areas inside the pores. The pore diameters must be wide enough to admit reactants
but narrow enough to exclude very large residue molecules.
In fixed-bed units, hydrotreating catalysts last for 1–5 years, typically 2–3 years.
During a catalyst cycle, the catalysts slowly deactivate and temperatures are raised to
compensate. Start-of-run reactor average temperatures can range from 600 to 700 °F
(315–370 °C), depending on hydrogen partial pressure, feed rate, feed quality, and
desired product quality. When the required temperature reaches a limit set by metal-
lurgy, the feed rate must be reduced, the product quality objective must be relaxed, or
the catalyst must be changed. Typical end-of-run average temperatures range from
780 to 800 °F (416–427 °C), and typically end-of-run peak temperatures range from
800 to 825 °F (427–440 °C). A catalyst cycle can end for other reasons, such as
a predetermined schedule, unacceptable pressure drop, or excessive production of
low-value light gases.
Spent catalysts are fouled with coke. They are either regenerating offsite by careful
combustion of the coke, or sold to catalyst reclamation companies. In-place regen-
eration is now very rare. Catalysts heavily contaminated with trace elements are not
regenerated. It is not unusual to see uncontaminated catalysts restored to 85–95% of
their initial activity.
Mercaptan oxidation processes, such as the two-step UOP Merox process, convert
foul-smelling mercaptans into disulfides. For Merox, the overall reaction is as fol-
lows:
4 R - SH + O2 → 2 R - S - S - R + 2 H2 O
2 R - SH + 2 NaOH → 2 NaSR + 2 H2 O
The next step regenerates the sodium hydroxide while producing a disulfide:
186 9 Crude Storage, Blending, Desalting, Distillation and Treating
4 NaSR + 2 O2 + H2 O → 2 R - S - S - R + 4 NaOH
In the flow diagram, shown in Fig. 9.21, the mercaptan-containing feedstock enters
the prewash vessel, which removes H2 S. A coalescer prevents caustic from being
carried out of the vessel. The feed then enters the extractor, flowing up through several
contact trays, which enhance the mixing of feedstock with caustic. The feed then goes
to a settler, in which the gas separates from the caustic. The gas is water-washed to
remove any residual caustic, and sent through a bed of rock salt to remove entrained
water. The dry, sweet LPG exits the Merox unit. The hydrocarbon-rich caustic from
the bottom of the extractor is mixed with a proprietary liquid catalyst, heated in
an exchanger, then mixed with compressed air. The mixture goes to the oxidizer,
where the extracted mercaptans are converted to into disulfides. The caustic/disulfide
mixture flows to a separator, in which the mixture separates into a disulfide layer
on top of an aqueous layer of “lean” caustic. The disulfides go to storage or to a
hydrotreater. The caustic is recycled back to the extractor.
References 187
References
10.1 Introduction
After Ford Model-T, there was a rapid growth in transportation vehicles. Motor
gasoline (“mogas”) became in high demand. Hence, many of early refining processes
were designed for conversion to increase gasoline yields. The carbon number range of
gasoline is between C5 and C12 . This fraction can be obtained as straight run naphtha
by distillation. However, not all naphtha molecules have good octane ratings suitable
for automobile performance. Isomerization and reforming into structures with high
octane ratings without altering the carbon numbers would be one approach. Other
approaches include combining C3 to C5 molecules by alkylation and polymerization,
or decreasing the size of C12 + molecules by cracking and coking, which will be
discussed in the following two chapters.
Normal paraffins of the lighter naphtha fraction, butanes, pentanes and hexanes, have
poor octane ratings for gasoline engines. Isomers of normal paraffins, on the other
hand, have high octane ratings. During World War II, there was a demand of high-
octane aviation gasoline, and isomerization became an important process. However,
the majority of butane isomerization units were shut down after World War II due
to lower demand. Demand was reduced mainly because tetraethyl lead (TEL) was
being added to gasoline to boost octane. TEL was banned after research revealed
environmental and health hazards. The phase-out of TEL created an octane gap,
which resulted in new butane isomerization units being installed, with AlCl3 as a
catalyst. Subsequently, n-pentane and n-hexane isomerization processes were devel-
oped, using supported acid catalysts that can stand high temperature (370–480 °C)
and high pressure (300–750 psi). A liquid-phase process employs a dissolved catalyst
for C5 /C6 isomerization.
because the isomerization reaction itself does not consume hydrogen. HCl is removed
in a stripper column. The liquids go to a stabilizer which separates C4 -(compounds
with carbon numbers less than 4), which can be used as a fuel gas inside refinery, from
isobutene and unconverted n-butane. The C4 compounds are recycled and mixed with
fresh feed. Product isobutane is collected at the top of deisobutanizer. Isobutane is the
key feedstock for alkylation units, which produce high octane alkylates, a valuable
component of gasoline.
Commercial processes have also been developed for the isomerization of low-
octane normal pentane and normal hexane to higher-octane isoparaffins (see below).
Here the reaction is usually catalyzed with supported platinum. As in catalytic reform-
ing, the reactions are carried out in the presence of hydrogen. Hydrogen is neither
produced nor consumed in the process but is employed to inhibit undesirable side
reactions. The reactor step is usually followed by molecular sieve extraction and dis-
tillation. Though this process is an attractive way to exclude low-octane components
from the gasoline blending pool, it does not produce a final product of sufficiently
high octane to contribute much to the manufacture of unleaded gasoline.
The C5 /C6 isomerization process is shown in Fig. 10.2. The mixed C5 /C6 feed
enters a fractionator to separate n-C5 /C6 from iso-C5 /C6 , which are the desired iso-
merization products. The purified n-C5 /C6 is then mixed with hydrogen and hydro-
gen chloride and heated to 240–500 °C under 300–1000 psi pressure before entering
the reactor, which is packed with HCl-promoted AlCl3 catalyst or a Pt-containing
catalyst. The residence time is 10–40 min. Again, hydrogen is used to control side
reactions (disproportionation and cracking). The product stream goes through a flash
tank to recover hydrogen for reuse, then enters a stabilizer to separate C4 and lighter
impurities before a C5 and C6 splitter. The C5 components are recycled back to the
feed fractionator to recover i-C5 into isomerate, a valuable component of gasoline.
A two-stage reactor scheme is shown in Fig. 10.3. The first reactor contains a
Pt-catalyst, which serves to hydrogenate any olefins and aromatics in the feed. It
also converts organic chloride into HCl removes any other impurities carried by the
feed stream. The second reactor is packed with AlCl3 which is promoted by HCl to
perform isomerization.
192 10 Gasoline Production
Fig. 10.3 C5 /C6 Isomerization with two reactors and once-through hydrogen
10.2 Isomerization 193
Figure 10.4 shows the UOP Penex isomerization unit with a similar design [1].
Penex uses a platinum-containing catalyst. The process improves octane ratings from
50–60 to 82–86 or higher.
In the Shell Hysomer process for pentane/hexane isomerization [3], the feed is
combined with hydrogen-rich gas, heated to 445–545 °F (230–285 °C) and routed to
the Hysomer reactor at 190–440 psi (13–30 bar). As with fixed-bed hydrotreating and
hydrocracking, the process fluids flow down through the catalyst bed, where a part
of the n-paraffins are converted into branched paraffins. The catalyst is comprised
of a noble metal on a zeolite-containing support. The reactions are exothermic, and
temperature rise is controlled by injecting relatively cold quench gas. The reactor
effluent is cooled by heat exchange and sent to a flash drum, which separates hydrogen
from the liquid product. The hydrogen is recycled. The liquid is fractionated, and the
n-paraffins are recycled. The net conversion of n-paraffins into branched products
can be as high as 97%, and the octane can be boosted by 8–10 numbers.
Often, the heat-exchanger and fractionation systems of isomerization units are
integrated with those of other process units, such as catalytic reformers. In the Union
Carbide total isomerization process (TIP) [4], C5 /C6 isomerization was integrated
with molecular sieve separation, which provided complete conversion of n-paraffins.
The CDTech Isomplus process that was jointly developed with Lyondell Petro-
chemical achieves near-equilibrium conversion of n-butenes into isobutylene, and n-
pentenes into isoamylene over a highly selective zeolite catalyst [5]. It was developed
when methyl t-butylether (MTBE) was in high demand for reformulated gasolines.
Isobutylene is also an important feedstock for polyisobutylenes.
Olefin isomerization converts straight-chain C4 –C6 olefins into corresponding
iso-olefins which can be used as alkylation feeds with excess an amount of isobutane
for producing high octane gasoline.
194 10 Gasoline Production
Straight-run gasoline from distillation has a very low octane number, i.e., it has a high
tendency to knock. Thermal reforming was initially developed from thermal cracking
processes with naphtha as the feed. The feed with an end point of 400 °F is heated to
950–1100 °F under 400–1000 psi pressure. The higher octane number is due to the
cracking of long-chain paraffins into high-octane (65–80) olefins. Gases, residual oil
or tar can be formed. Thermal reforming is less effective and less economical than
catalytic reforming. Catalytic reforming was commercialized during the 1950s to
produce reformate with research octane number on the order of 90–95 [6]. Catalytic
reforming now furnishes approx. 30–40% of US gasoline requirements.
Catalytic reforming is conducted in the presence of hydrogen over hydrogena-
tion-dehydrogenation catalysts supported on alumina or silica-alumina. The oper-
ation conditions for modern catalytic reforming are: temperature at 840–965 °F,
pressure at 100–600 psi. Continuous catalyst regeneration (CCR) reformers operate
at the low end of the pressure range, 100–150 psig, where production of aromat-
ics is more favored by thermodynamics. The first reforming catalysts were based
on molybdena-alumina or chromia-alumina, but since the commercialization of the
Platforming process in 1949, noble metal catalysts have been used, either with sup-
ported platinum alone or with supported platinum-rhenium, platinum-rhenium-tin,
or other tri-metallics on a silica-alumina or alumina base [7]. Hydrogen chloride is
a co-catalyst.
Naphtha containing seven or more carbons is used as a reformer feed. Pentane and
hexane are removed from desulfurized naphtha to be used as isomerization feeds,
as discussed above. If the feed contains C6 hydrocarbons, carcinogenic benzene is
hydrogenated into cyclohexane to avoid its release into atmosphere upon incomplete
combustion or when drivers fill their automobile tanks with gasoline. There are four
main hydrocarbon types in naphtha: paraffins, olefins (mainly from catalytic crack-
ing and coking units), naphthenes, and aromatics. Organic sulfur and most olefins
are removed by prior hydrotreating; the sulfur content must be <1 ppmw. During the
process, paraffins are cyclized into naphthenes, and naphthenes undergo dehydro-
genation to form aromatics. Aromatics are left essentially unchanged. The preferred
conditions for dehydrogenation are: high temperature, low pressure, and low space
velocity.
An important co-product is hydrogen, which is required by other processes, espe-
cially hydrotreating and hydrocracking. The dehydrogenation reaction in reforming
Table 10.1 Typical feed and Component Feed (%) Product (%)
product distributions in
reforming Paraffins 30–70 30–50
Naphthenes 20–60 0–3
Aromatics 7–20 45–60
10.3 Catalytic Reforming 195
methylcyclopentane
Acid site
C+ intermediate CH3
Alky shift
CH3CH2CH2CH2CH2CH3 CH3CCH2CH2CH3
CH3 - H2 - 3 H2
n-hexane 2-methylpentane benzene
is endothermic. Heat is required. Table 10.1 lists typical feed and product composi-
tions in a reforming process.
Figure 10.5 summarizes the reactions of catalytic reforming, the purpose of which
is to transform C6 to C11 naphthenes and paraffins into aromatic compounds. The
aromatics can go to chemical plants or be used as high-octane gasoline blend stocks.
For the paraffins in the feed, the acidic sites of the catalyst isomerize n-paraffins to
iso-paraffins. This is a key step for cyclization. It is followed by dehydrogenation to
form aromatic compounds. The process yields considerable amounts of hydrogen;
in Fig. 10.5, four moles of hydrogen are produced from one mole of hexane. The
hydrogen is used in hydrotreaters, hydrocrackers, isomerization units, and others.
In contrast to hydrocracking, which operates at high pressure, catalytic reforming
operates at low pressure and high temperature, which favors production of aromat-
ics. The isomerization of alkylcyclopentanes into cyclohexanes is a key step in the
production of benzene and alkylbenzenes.
The three major process flows for catalytic reforming are:
• Semi-regenerative
• Cyclic (fully regenerative)
• Continuous catalyst regeneration (CCR, moving bed).
Figure 10.6 shows a reactor for a fixed-bed semi-regenerative reformer. The feed
is introduced through the top and flows down along the outer wall of the reactor and
then flows radially inward through the thin annular catalyst bed. Screens are used
to contain the catalyst and allow reactants to enter from the wall, pass through the
catalyst bed, enter the central collection tube and exit the vessel at the bottom. This
design allows the intimate contact between the feed and the catalyst.
The process flow is shown in Fig. 10.7. Catalyst cycles last from 6 to 12 months.
A cycle ends when the unit is unable to meet its process objectives—typically C5 -
plus yields >80 wt% when the octane number target is 100 RON. At the end of
a cycle, the entire unit is brought down and coke is burned off the catalyst. The
depentanized/dehexanized naphtha is hydrotreated to bring sulfur down to <5 ppm,
then mixed with hydrogen and heated to >900 °F (>480 °C). The hot mixture passes
196 10 Gasoline Production
through a series of fixed-bed reactors. The feed is spiked with an organic chloride,
which converts to hydrogen chloride in the reactors. This provides the required
catalyst acidity and helps minimize catalyst coking. There are three reactors shown.
10.3 Catalytic Reforming 197
Naphtha Feed
In alkylation, isobutane reacts with C3 to C5 olefins, which come mainly from crack-
ing units, in the presence of strong acids to produce branched chain hydrocarbons.
Without the presence of an excess amount of isobutane, C3 to C5 olefins can undergo
undesired polymerization into sludge. These hydrocarbons, often referred to as alky-
late, have high octane values (motor octane number from 88 to 95 and research octane
number from 90 to 98) and low Reid vapor pressures, making them an excellent con-
tributor to the gasoline blending pool.
Figure 10.9 shows the mechanism for the alkylation reaction of butenes with
isobutane [1, 14]. The process uses an excess of isobutane to control heat and mini-
mize olefin polymerization. As an initiation step exemplified in Reaction 1, 2-butene
or its isomer is protonated by the acid catalyst to form a 2-butenium ion, which
isomerizes to stable t-butyl ion. Reactions 2a, 2b and 2c show how different isomers
lead to different C8 carbocations. These C8 carbocations will isomerize into more
stable 2,2,4-trimethyl pentyl ion. The isomerization reactions will not reach com-
pletion due to thermodynamic equilibrium of reversible reactions, evidenced by the
presence of small amounts of 2,2,3-trimethyl pentane and 2,2-dimethyl hexane in the
product. The C8 carbocations ions will react with predominant amount of isobutane
through hydride transfer to be neutralized and leave t-butyl ion for propagating the
alkylation reaction. However, some C8 —carbocations (C4 dimer ions) can proceed
to react with olefins to form C12 -carbocations (C4 trimer ion) and continue poly-
10.4 Alkylation 199
hydrocarbons, including isobutane. Olefins are not present in crude oils. Hence, they
cannot come from distillation directly. They are mainly from thermal conversion units
in the refinery, such as thermal cracking, catalytic cracking, hydrocracking, coking,
etc. The main source of olefins, particularly C=3 and C=4 , is from a fluid catalytic
cracking (FCC) unit, shown in the figure.
The flow inside an alkylation unit is shown in Fig. 10.12. The alkylation reaction is
highly exothermic, so the mixed feed of isobutene and cracked gas containing olefins
is chilled to a low temperature before entering the reactor. The low temperature also
decreases polymerization of the olefins. After the reaction, the acid is recovered and
recycled. The product is caustic washed to remove the residual acid from the product.
Then the product stream goes through the depropanizer to separate propane, to the
deisobutanizer to recover isobutane for reuse, and to the debutanizer to separate
butane from the desired product, alkylate.
Alkylate is an ideal gasoline component. It has a octane numbers exceeding 90.
During World War II, alkylation became the main process for improving the octane of
aviation gasoline (avgas). Alkylate has zero sulfur, zero benzene, and zero olefin con-
tent, so it is friendly to the environment. Its low Reid Vapor Pressure (RVP) reduces
hydrocarbon emissions. In contrast, blending n-butane into gasoline increases RVP.
With 10% ethanol in reformulated gasoline, RVP is higher by 1.1 to 1.25 psi [15].
Hence, there is no more need to add n-butane during winter or very cold weather for
startup of gasoline engines. This causes a seasonal glut of n-butane.
Alkylation employs either hydrofluoric acid, as in a Hydrofluoric Acid Alkylation
Unit (HFAU), or sulfuric acid, as in a Sulfuric Acid Alkylation Unit (SAAU). The
process usually runs at low temperatures to avoid polymerization of the olefins.
Temperatures for HF catalyzed reactions are 70–100 °F (21–38 °C). For sulfuric
acid they are 35–50 °F (2–10 °C), so SAAU alkylation requires feed refrigeration.
10.4 Alkylation 201
Fig. 10.11 Refinery flow with alkylation unit (APS: atmospheric pipe still or atmospheric distil-
lation unit, VPS: vacuum pipe still or vacuum distillation unit: SR: straight run; FCC: fluidized
catalytic cracking unit)
The process variables affect the product quality in octane number and the yield.
In the HFAU, increasing temperature from 60 to 125 °F (16–52 °C) will decrease the
octane number about 3. In the SAAU, increasing temperature from 25 to 55 °F (−4
to 13 °C) decreases octane number from one to three. In HFAU, acid concentration
is in the range of 86 to 90 wt%, while in the SAAU, between 93 and 95%. The
isobutane:olefins ratio is between 5:1 and 15:1. The olefin space velocity, which is
defined as the volume of olefins divided by the volume of the acid in the reactor, is
4–25 min. contact time in HFAU and 5–40 min. in SAAU. In general, lowering the
olefin space velocity reduces the amount of high-boiling hydrocarbons produced,
increases the product octane number, and reduces acid consumption. Mrstik, Smith
and Pinkerton developed a correlation factor to predict alkylate quality with operating
variables, defined as:[16]
IE (I O)F
F=
100(SV)O
Figure 10.15 shows an autorefrigeration sulfuric acid alklylation unit. The olefin feed
is mixed with isobutane and acid, and chilled to 35–45 °F through a heat exchanger
for minimizing redox reaction and preventing tar and SO2 formation. A pressure at
5–15 psi is applied to prevent vaporization. Propane and lighter gases are withdrawn
at the top of the reactor. The gases are compressed and liquefied. A portion of this
liquid is vaporized in an economizer to cool the olefin feed, as autorefrigeration,
before it is sent to the reactor or a depropanizer.
The C4 ’s and alkylate mixture is sent to deisobutanizer, after caustic wash to
remove residual acid, to fractionate into isobutane for recycle with the isobutane
feed, n-butane and alkylate.
A typical Stratco effluent refrigerated sulfuric acid alkylation unit is shown in
Fig. 10.16 which uses acid settler effluent as the refrigerant to provide cooling in
the contactor reactor. The Stratco contactor reactor is a horizontal pressure vessel
containing an inner circulation tube, a tube bundle to remove the heat of reaction,
and a mixing impeller. Hydrocarbon feed and recycle acid enter on the suction side
of the impeller inside the circulation tube. As the feeds pass across the impeller, a
fine emulsion of hydrocarbon and acid is formed by the extremely high shear forces
induced by the impeller. Heat transfer from the reaction side of the tube bundle to
the refrigeration side is aided by the high circulation rate, which also prevents any
significant temperature differential within the contactor reactor. A portion of the
circulating emulsion in the contactor reactor flows from the circulation tube, on the
discharge side of the impeller, to the acid settler, where the hydrocarbon phase is
separated from the circulating acid phase. The acid is recycled back to the contactor.
Hydrocarbon leaves the settlers and is let down across a back-pressure control valve
to the tube side of the contactor reactor bundle. The heat of reaction from the shell side
204 10 Gasoline Production
Currently, more than half of the world’s approximately 700 refineries have alkylation
units that use HF or H2 SO4. HF alkylation process has several advantages over
sulfuric acid alkylation process: (1) the reactors are smaller and simpler; (2) the HF
process is less sensitive to temperature than the H2 SO4 process,cooling water can be
used instead of refrigeration; (3) it has a smaller settling device for emulsions; (4)
10.4 Alkylation 205
stream is neutralized off-site; (3) drying is not required for H2 SO4 process, while
drying down to a few ppm water is needed for the HF process. In the H2 SO4 process,
only feed coalescers are used to remove the free water that drops out of the chilled
feed; (3) additional safety equipment is required for HF process, with greater costs;
(4) capital costs for HF process are higher; (5) self-alkylation of isobutane occurs in
the HF process; and (6) H2 SO4 alkylation works better for butylene.
However, due to their strong acidity in the liquid phase, both HF and H2 SO4 require
relatively expensive corrosion-resistant vessels and equipment. Safety in handling
and operations is a major concern for both, which also face disposal issues associated
with spent acids and acid-soluble oils (ASO). These problems can be diminished with
ionic-liquid alkylation and essentially eliminated by solid acid-catalyzed alkylation.
These recently developed breakthrough processes are discussed below.
Strongly acidic ionic liquid catalysts are safer and easier to handle than HF and
H2 SO4 . The ionic liquid catalyst can be used in a lower volume at temperatures
below 100 °C. The catalyst can be regenerated on-site, giving it a lower environmental
footprint than HF and H2 SO4 technologies.
The world’s first composite ionic liquid alkylation (CILA) commercial unit of
120 kt/a was commissioned at Shandong Deyang Petrochemical Plant in Dongy-
ing, China in August, 2013, after a successful pilot plant test run. The process
was retrofitted into an existing 65 kt/a sulfuric acid alkylation unit at a PetroChina
Lanzhou Petrochemical plant starting in 2005, using the technology developed by
China University of Petroleum [20]. The alkylation of isobutylene occurs in a strongly
Lewis acidic ionic liquid based on aluminum (III) chloride. The retrofit of the existing
sulfuric acid alkylation unit, called ionikylation, not only increased the yield of the
process (compared to sulfuric acid), but also increased the process unit capacity by
40%, with attractive economics. This is by far the largest commercial usage of ionic
liquids reported to date [21].
In the U.S., Honeywell UOP announced in 2016 the commercialization of
ISOALKYTM , a new alkylation technology developed by Chevron USA [22].
Chevron proved the technology in a small demonstration unit at its Salt Lake City
refinery, where it operated for five years. Due to the success of the small unit, Chevron
committed to convert its hydrofluoric acid (HF) alkylation unit in Salt Lake City to
ISOALKY technology. The completed ISOALKY commercial unit will be opera-
tional in 2020 [23].
The catalyst for the process is a highly acidic ionic liquid—a non-aqueous liquid
salt. The process operates at temperatures below 100 °C. The net reactions are the
same as for other alkylation processes: isobutane + C3 –C5 olefins → alkylate. The
ionic liquid catalyst performs as well or better than HF and H2 SO4 , but with lower
volatility. Due to the lower vapor pressure, the ionic liquid is easier to handle. HF
alkylation units can be cost-effectively converted to ISOALKY technology. Other
10.4 Alkylation 207
advantages include the ability to produce alkylate from a wider range of feedstocks
using a lower volume of catalyst. The catalyst can be regenerated on-site, giving it a
lower environmental footprint than HF and H2 SO4 technologies.
Olefin gases can be polymerized to liquid products that are suitable for gasoline in
the carbon number range between 5 and 12. The term “polymerization” has a unique
meaning in the gasoline industry. It is not the same as making high polymers, such
as polyethylene and polypropylene. A better name for gasoline “polymerization” is
“oligomerization” because of the low number of repeating olefin units in the reaction.
Polymer gasoline (poly gasoline or polymerate) exhibits high octane values. A typical
reaction is shown in Fig. 10.17. The usual feedstock is propylene and butylene from
cracking processes. Catalytic polymerization came into use in the 1930s and was one
of the first catalytic processes to be used in the petroleum industry.
Figure 10.18 shows a simplified diagram for the polymerization of olefin unit. The
feed is propylene and butylenes from thermal and catalytic cracking units. Sulfur- and
oxygen-containing compounds are removed from the feed. Amine treating removes
H2 S and caustic washing removes mercaptans. The gas is then passed through a
scrubber in which H2 O to removes caustic or amines. A small amount of H2 O is added
to compensate for water loss during the polymerization, thus maintaining catalyst
activity. The catalysts can be sulfuric acid, copper pyrophosphate, solid phosphoric
acid (SPA) on pellets of kieselguhr (a porous sedimentary rock), or some similar
material. The reaction conditions are temperature at 150–220 °C (300–425 °F) and
pressure at 150–1200 psi (10–80 atm). The reaction is exothermic. Cold recycled
propane is used as a coolant to quench the reaction. The product mixture is sent
to a flasher to separate propylene and butylene for recycle, while the bottom enters
a stabilizer to separate C3 /C4 from polymer gasoline product. Polymer gasolines
derived from propylene and butylene have octane numbers above 90.
Figure 10.19 shows a flow diagram for the UOP SPA polymerization process. The
SPA catalyst is a combination of Kieselguhr and phosphoric acid [26]. The conditions
can be varied to produce maximum polymer yield at almost complete conversion of
the olefins. In the polymerization of butylenes the process can be carefully controlled
to yield largely isooctenes, which on hydrogenation will yield isooctanes having an
References
1. Hsu CS, Robinson PR (2017) Gasoline production and blending. In: Hsu CS, Robinson PR
(eds) Springer handbook of petroleum technology, New York, Springer
2. Dean LE, Harris HR, Belden DH, Haensel V (1959) The Penex process for pentane isomer-
ization. Platimum Metal Rev. 3(1):9–11
3. The petroleum handbook: edition 6 by shell, Elsevier (1986)
4. Holcombe TC (1980) Total Isomerization Process, US Patent 4210771A, Jul 1, 1980
5. Zak T, Behkish A, Shum W, Wang S, Candela L, Ruszkay J (2009) “Isomerization of Butenes:
LyondellBasell’s isomplus technology developments”, presented at Production and Use of
Light Olefins, DGMK Conference 28–30 Sept, Dresden, Germany
6. Antos GJ, Aitani AM (eds) (2007) Catalytic naphtha reforming. 2nd edn. New York, Marcel
Dekker
7. Hsu CS, Robinson PR (2017) Gasoline production and blending. In: Hsu CS, Robinson PR
(eds) Springer handbook of petroleum technology, New York, Springer
8. LeGoff P-Y, Kostka W, Ross J (2017) Catalytic reforming for fuels and aromatics, In: Hsu CS,
Robinson PR (eds) Springer handbook of petroleum technology, Heidelberg, Springer
210 10 Gasoline Production
Thermal cracking was first invented and patented by a Russian engineer, Valdimir
Shukhov, in 1891. However, its development was not pursued beyond the laboratory.
American engineers, William Merriam Burton and Robert E. Humphreys indepen-
dently developed a similar process in 1908, but Burton filed for a patent in 1912; the
patent, which is of great significance in the history petroleum refining, was granted
in 1913 (U.S. 1,049,667). The first battery of twelve stills used in thermal cracking
went into operations at the Witting Refinery of Standard Oil of Indiana (now BP) in
1913. Early versions were batch processes. However continuous processes eventu-
ally took over in the 1920s. Once catalytic cracking was invented in the early 1940s,
it became much more popular than thermal cracking. More sophisticated forms of
thermal cracking have been developed for various purposes, including steam crack-
ing, visbreaking and coking. Modern high-pressure thermal crackers, which operate
at absolute pressures of about 7000 kPa, are the basis for economically important
production of olefins for polymers.
Thermal cracking is similar to what occurs below the surface of the Earth when
kerogen is broken down into lighter components. The actual reaction mechanism
behind cracking is very complex, and computers model hundreds or thousands of
reactions for the process, but the main reactions are all related to free radical chem-
© Springer Nature Switzerland AG 2019 211
C. S. Hsu and P. R. Robinson, Petroleum Science and Technology,
https://doi.org/10.1007/978-3-030-16275-7_11
212 11 Cracking
RCH2CH3 + H• RCH2CH2• + H2
CH3CH3 2 CH3•
RCH2CH2• + CH4 RCH2CH3 + CH3•
Radical Addition
CH3
Radical Decomposition
RCH=CH2 + •CH3 RCHCH2•
RCH2CH2• H2C=CH2 + R•
CH3 CH3
CH3
RCHCH2• + R’CH=CH2 RCHCH2R’CH2CH2•
RCHCH2• H2CCH=CH2 + R•
+ + 4 H2
istry. At the high temperatures, free radicals are more likely to form. A free radical
is a molecule with an unshared electron attached to it. This unshared electron can
break other bonds leading to a series of chain reactions. Each subsequent reaction will
either create another free radical, which can react further, or combine with another
free radical to end the chain.
Figure 11.1 outlines mechanisms postulated by Greensfelder, et al [1]. for the
thermal cracking of hydrocarbons. This chain-reaction mechanism includes the fol-
lowing steps:
(1) Chain Initiation: Radicals—neutral atoms or compounds with a free electron
and no charge—are formed due to direct thermal rupture of a chemical bond.
Common initiators (other than heat) are organic sulfur and oxygen compounds,
which are present in delayed coker feeds. H2 O2 and Cl2 are potent initiators.
In steam cracker feed, heteroatom initiators are absent, so the chain reaction is
initiated by the rupture of a C–C bond at very high temperature.
(2) Propagation: A radical abstracts hydrogen from another compound, turning it
into a different radical.
(3) Radical addition: A radical attacks an olefin group, forming a C–C bond and a
larger radical.
(4) Isomerization: A primary radical isomerizes into a more stable secondary radi-
cal.
11.1 Thermal Cracking 213
Steam cracking, shown in Fig. 11.3, is a process of breaking down saturated hydro-
carbons into smaller, often unsaturated, hydrocarbons by reactions with steam in a
bank of pyrolysis furnaces. It is the principal industrial method employed at olefin
plants for producing ethylene and propylene, important feedstocks for polyolefins
(polyethylenes, polypropylenes, etc.), which account for 50–60% of all commer-
cial organic chemicals. Propylene is also produced by other methods, such as propane
dehydrogenation [2]. Fluid catalytic cracking (FCC) is an important commercial
source of propylene and butylene. Steam cracking in olefin plants co-produces hydro-
gen, which is often used in refineries.
The feed can be gaseous or liquid hydrocarbons, such as ethane, LPG (propane
and butane), naphtha, and unconverted oil from hydrocracking units. C4 olefins,
including butadiene, and benzene-rich pyrolysis gasoline are produced when heavy
liquid feeds, such as naphtha and hydrocracker gas oils, are used. The feeds are
diluted with steam and briefly heated to 1050 °C in a furnace without the presence of
oxygen and fed to Cr–Ni reactor tubes. The hydrocarbon chains preferentially crack
at the center of molecules at 400 °C. Cracking shifts toward the end of the molecule
with increasing temperature, leading to larger quantities of preferred low molecular
weight olefin products. Thus, the reaction temperature is very high, with a reactor
outlet temperature around 850 °C. The residence time is very short to improve yield
and to avoid coke formation and oligomerization. In modern furnaces, reaction time
can be reduced to milliseconds, with velocities faster than the speed of sound. The
reaction is stopped by rapid quenching to 300 °C. The product mixture is scrubbed
to remove H2 S and CO2 and then dried before it is sent to a series of separation
columns to separate and recover methane, ethane, ethane, propane, propylene and
C4 hydrocarbons. Figure 11.4 shows a steam cracking unit outside London with tall
11.2 Steam Cracking 215
separation columns in series for the production and separation of light hydrocarbon
gases.
Steam cracking product distribution depends on the feed composition, hydrocar-
bon to steam ratio, cracking temperature and residence time in the furnace. Light
hydrocarbon gas feeds, such as natural gas, ethane or LPG yield product streams
rich in the light olefins, essentially ethylene and propylene. In addition to light
olefins, heavier liquid feedstocks can also produce butylenes, butadiene and prod-
ucts rich in aromatic hydrocarbons suitable for blending into gasoline and fuel oil,
or routed through an extraction process to recover BTX aromatics (benzene, toluene
and xylenes).
Thermal cracking was used prior to 1925 for producing naphtha through thermal
decomposition of larger molecules into smaller molecules. In the late 1920s, Eugene
Houdry demonstrated that a catalytic cracking process yielded more gasoline of
higher octane-number, with less heavy fuel oils and light gases. Thermally cracked
216 11 Cracking
naphtha is quite olefinic, while catalytically cracked naphtha contains fewer olefins
and large amounts of aromatics and branched compounds. The first commercial
fixed bed cat cracking unit began production in 1937. The catalysts are covered by
a deactivating layer of coke in a short time during the process. The catalysts can
be regenerated by burning off the coke, but the time is relatively slow compared to
reaction time. It is more efficient to move the catalyst from one reactor for cracking
hydrocarbons to another reactor for catalyst regeneration.
In refining, catalytic cracking falls into two categories: catalytic cracking in the
absence of external hydrogen—primarily fluid catalytic cracking (FCC)—and cat-
alytic cracking in the presence of external hydrogen—hydrocracking.
Catalysts for both kinds of catalytic cracking contain strong acid sites. The mecha-
nism involves carbocations, also known as carbenium ions or carbonium ions. Acidity
is provided by amorphous silica/alumina or a crystalline zeolite. Commercial cata-
lysts for these and other processes are discussed in a subsequent section.
FCC units produce aromatics and other heavy products via cyclization, alkyla-
tion and polymerization of intermediate olefins. Polyaromatics can grow into larger
polyaromatics, which eventually can form coke.
Key features of catalytic cracking chemistry include a preponderance of branched-
chain paraffin products and low yields of methane and ethane. If there is a deficiency
of hydrogen, such as in the FCC process, hydrogen is produced and significant
amounts of small olefins are formed.
These days, most refiners pretreat FCC feeds in a fixed-bed hydrotreater. The
hydrotreater removes trace metal contaminants such as nickel and vanadium. Oth-
erwise, nickel would increase coke formation and decrease liquid yields. Vanadium
reduces conversion, decreases liquid yields, and destroys the catalyst. In addition to
removing Ni and V, the pretreater decreases concentrations of sulfur, nitrogen, and
aromatics. In the FCC regenerator, sulfur on the coked catalyst is converted to sulfur
oxides (SOx) in the flue gas. Clean air regulations restrict SOx emissions, which
cause acid rain. Therefore, removing sulfur from the FCC feed—thereby reducing
SOx formation—is highly beneficial. Removing nitrogen is beneficial, too, because
basic feed nitrogen suppresses the activity of highly acidic FCC catalysts. Pretreating
also saturates aromatics. As we have seen, saturating aromatics makes them more
crackable, so pretreating increases FCC conversion, often by more than 10 vol.%.
The underlying mechanism for catalytic cracking is essentially the same for both fluid
catalytic cracking and hydrocracking. Process differences are due to differences in
11.3 Catalytic Cracking 217
catalysts, equipment, and operating conditions. FCC produces a high yield of naph-
tha (gasoline) and LPG. It is considered as the heart and workhorse of a fuels refin-
ery. FCC produces more than half the world’s gasoline. It generates middle distil-
late streams (cycle oils) for further refining or blending. It also produces, via heat
exchange, a large quantity of high-quality steam.
FCC chemistry is a complex mixture of many reactions. The strong acid sites, both
Lewis and Brønsted, are supplied by zeolites, which are key components of modern
FCC catalysts. Influenced by their unique pore geometry, zeolites (usually H–Y) are
far more efficient at generating gasoline and middle distillate products than either
thermal cracking or cracking catalyzed by amorphous silica-alumina. A comparison
of zeolite and amorphous Al–Si cracking is shown in Table 11.1.
There are several possible initiation steps, shown in Fig. 11.5 [7]. One type of initia-
tion step (Reaction 11.1) is mild thermal cracking to generate free radicals, followed
by radical recombination to generate olefins. FCC reactions are mainly catalytic as
evidenced by the fact that they produce relatively small amounts of methane, the
generation of which is significant in thermal cracking. Methane can be formed by
hydrogen abstraction by a methyl radical or by combining a methyl radical with
hydrogen radical; the latter is a termination step of free-radical reactions. The main
initial steps involve Brønsted acid protonation of olefins to generate secondary (2°)
carbocations, shown in Reaction 2, and hydride abstraction from alkanes at Lewis
base catalyst sites, shown in Reaction 3, also to generate (2°) carbocations. Of minor
importance is the reversible dehydrogenation of alkanes at metal-containing catalyst
sites.
Figure 11.6 shows how 2° carbocations rearrange. There can be 2°–2° rearrangement
and 2°–3° (tertiary) rearrangement; 3° carbocations are more stable.
218 11 Cracking
Thermal
cracking
R CH2 CH2 CH2 CH3 R CH2 CH2 CH2 + CH3
1
Brønsted Lewis
H+ addition H- removal 3
2 acid site acid site H2 / +H2 4 Metal Site
R CH=CH CH2 R’
+ +
R CH2 CH CH2 R’ R CH2 CH CH2 R’ + LH
2° carbocation 2° carbocation
H
|
| + +
R C C H R CH CH3
| |
| | 2° carbocation
H H
1° carbocation
H
|
| + +
R C C CH2 R C CH2 CH2
| | |
| | |
CH3 H CH3
2° carbocation 3° carbocation
+ +
CH3 C CH2 CH2 R CH3 CH2 C CH2 R
| |
H H
2° to 2° skeletal isomerization
+ +
(1) CH3CHCH2CH2CH3 X CH3CH=CH2 + CH2 CH3 1° carbocation produced:
less favored
CH3 CH3
| + |
(2) CH3 C CH2 CH CH3 CH3 CH=CH2 + CH3 C+ 3° carbocation produced:
| | favored
CH3 scission CH3
+
+
CH3 CH2 CH2 CH CH3 + RH CH3 CH2 CH2 CH2 CH3 + R
hydride transfer
Fig. 11.7 Catalytic cracking mechanism: beta scission and hydride transfer
Beta scission involves cleavage of a C–C bond in the position beta to the carbon
atom that carries the positive charge. It is the primary cracking reaction in FCC. It is
endothermic and favored at higher temperatures; however, if the temperature is too
high, thermal cracking can become significant. Beta scission, shown in Fig. 11.7, is
characteristic of mechanisms with carbocation intermediates.
Hydride transfer also is shown in Fig. 11.6. If hydride transfer is substantial, paraf-
fins (normal and branched) will be the predominant products: FCC catalysts can be
designed for greater or lesser hydrogen transfer.
H • +H• → H2
H • +R• → RH
R • +R • → R − R
220 11 Cracking
In FCC, a full range of smaller molecules is formed from the breakup of large
molecules. Due to the overall deficiency of hydrogen, significant amounts of olefins
are formed. Since the feed contains large aromatic and naphthenic molecules with
long side chains attached, side chain cleavage, which can also initiate the chain of
reactions, is common. The molecules from which side chains have been removed
have higher specific gravity, i.e., lower API gravity.
Cracking also generates coke via aromatic ring growth by successive cyclization,
polymerization and dehydrogenation. The coke coats the catalyst, rendering it inac-
tive. Hence, the catalyst needs to be regenerated by burning off the coke in order to
restore activity.
To allow onstream catalyst regeneration, a moving-bed unit, Thermofor catalytic
cracking (TCC), was developed in 1941 with catalyst cycled between the reactor and
regenerator. It has become obsolete and replaced by more advanced fluidized bed
units with greatly improved catalysts.
A simplified diagram of the TCC process is shown in Fig. 11.8 [8]. Its essential
elements are a reactor for continuously contacting hydrocarbon feed with a moving
bed of granular catalyst for conversion of the hydrocarbons, and a kiln (regenera-
tor) for removing carbon deposit from the catalyst during the cracking operation by
controlled combustion with air. The catalyst flows by gravity through both vessels,
which stand side by side, and is transferred from the bottom of one to the top of
the other by means of bucket elevators. All required process heat is supplied by
the highly exothermic combustion of coke in the regenerator. Catalyst temperatures
reach 1200–1500 °F. In contrast, the cracking reaction is endothermic. The feed is
preheated by heat exchange to 500–800 °F before entering the reactor. There, addi-
tional heat is supplied by mixing with the hot recycled catalyst to reach a temperature
of 900–1050 °F. The reactor outlet vapor is sent to a fractionator, where it is separated
into different fractions—gas (offgas), cracked naphtha, fuel oil, and slurry oil, which
can be mixed with fresh feed and recycled back to the reactor for further processing.
A typical FCC unit comprises three major sections—riser/reactor, regenerator,
and disengaging vessel. In the riser/reaction section, preheated oil is mixed with hot,
regenerated catalyst. The mixture acts as a fluid because the catalyst particles are
about the size of sifted flour. The hot catalyst vaporizes the oil, and the vaporized oil
carries the catalyst up the riser/reactor. The cracking reaction is very fast, achieving
completion in just a few seconds or even less. It produces light gases, high-octane
gasoline, and heavier products called light cycle oil (LCO), heavy cycle oil (HCO),
slurry oil, and decant oil. It also leaves a layer of coke on the catalyst particles,
rendering them inactive.
There are two basic types of FCC units: “side-by-side,” where the reactor and the
regenerator adjacent to each other, shown in Fig. 11.9, and the stacked type, in which
the reactor is mounted on top of the regenerator, shown in Fig. 11.10.
At the top of the riser, the riser outlet temperature (ROT) can reach 900–1020 °F
(482–549 °C). The ROT determines conversion and affects product selectivity, so
11.3 Catalytic Cracking 221
Bucket elevators
Regenerated
Catalyst
hopper
Reactor Regenerator
Feed in
Air in
fractionator to separate product components is also shown. The top effluent contains
light gases that are sent to a recovery system to remove sour gases for recovery of
flue gas, LPG, light olefins, etc. The bottom stream can be recycled by mixing with
feed for further processing.
FCC units must be heat-balanced, or they won’t run. The burning of coke deposited
on the catalyst in the regenerator provides all of the heat required by the process. In
fact, FCC units are significant sources of high-quality steam for other refinery units.
Table 11.2 gives a representative breakdown of FCC heat requirements.
Residue FCC (RFCC) units, also known as heavy oil crackers, can process significant
amounts of atmospheric residue (650 °F+) and vacuum residue to produce gasoline
and lighter components [9, 10]. It generates substantially more coke than conven-
tional FCC feeds. Excess heat is generated when the extra coke is burned in the
regenerator, and residues contain high amounts of trace metals, particularly nickel
and vanadium. Those metals destroy FCC catalysts. Residue FCC units must handle
both challenges.
The metals are removed in upstream hydrodemetalation (HDM) units [11]. Cat-
alyst coolers and supplemental regenerators recover the excess heat. The catalyst
coolers (e.g., steam coils) are installed usually on the second-stage regenerator. The
UOP catalyst cooler is an external vertical shell-and-tube heat exchanger [12]. Cat-
alyst flows across the tube bundle in the dense phase. UOP’s air lance distribution
system ensures uniform air distribution within the tube bundle and a uniform heat
transfer coefficient. The generation of steam (up to 850 psig) from the circulating
water is used to remove heat from the regenerated catalyst.
Three different styles of catalyst coolers—flow-through, back-mix and
hybrid—have been designed and commercialized to accommodate a wide range
of heat removal duties as well as physical and plot-space constraints [13].
Conventional hydrotreating does a good job of removing sulfur from FCC feed,
which leads to lower sulfur in FCC products. Unfortunately, despite pretreatment,
FCC gasoline can still contain up to 150 ppmw sulfur—far more than the present
specification of 10 ppmw. Hydrotreating removes sulfur from the gasoline, but it also
reduces octane by saturating C6 –C10 olefins.
In processes such as Prime-G+ [14], offered for license by Axens, full-range
naphtha is split into light and heavy fractions. The light fraction contains most of
the high-octane olefins but not much of the sulfur. After diolefins are removed via
selective hydrogenation, the light cut is ready for gasoline blending. The heavy
fraction contains most of the sulfur but not much of the olefins. It is hydrotreated
conventionally.
The S-Zorb process [15], invented by ConocoPhillips, uses selective adsorption
to remove sulfur from FCC gasoline. The feed is combined with a small amount
of hydrogen, heated, and injected into an expanded fluid-bed reactor, where a pro-
prietary sorbent removes sulfur from the feed. A disengaging zone in the reactor
removes suspended sorbent from the vapor, which exits the reactor as a low-sulfur
stock suitable for gasoline blending. The sorbent is withdrawn continuously from
the reactor and sent to the regenerator section, where the sulfur is removed as SO2
and sent to a sulfur recovery unit. The clean sorbent is reconditioned and returned to
the reactor. The rate of sorbent circulation is controlled to help maintain the desired
sulfur concentration in the product.
In practice, catalysts do change as time goes by and more materials are processed.
They degrade due to coking, attrition, feed contamination and/or process upsets.
Some catalysts last for years before they have to be replaced. In other processes,
such as FCC and CCR catalytic reforming, they are regenerated and reused inside the
process during normal operation. In the FCC process, degraded catalyst is removed
and replaced with fresh catalyst as needed. The chemistry of coke formation on
catalysts is similar to the coke formation mechanism presented in Chap. 9, Sect. 4.
Catalysts facilitate reactions by decreasing activation energies. Consider ammonia
synthesis:
N2 + 3H2 → 3NH3
Fig. 11.11 Structures of zeolites: ZSM-5 (a), mordenite (b), beta (c), MCM-22 (d), zeolite Y (e),
and zeolite L (f)
Fig. 11.12 Brønsted (B) and Lewis acid sites (L) in zeolites and amorphous silica/aluminas
ton donors). The H+ can be replaced by other positive counter-ions such as Na+ , K+ ,
and NH4 + . The counter ions can be swapped via ion exchange. For example, when
Na-Y zeolite is exchanged with an ammonium salt, the Na+ ion is replaced by NH4 + .
When NH4 -Y is heated to the right temperature, the ammonium ion decomposes,
releasing NH3 (gas) and leaving behind highly acid H-Y zeolite. A typical Si/Al ratio
for a zeolite with high cracking activity is >5. In hydrocracking catalysts with less
activity but greater selectivity for production of middle distillates, the Si/Al might
be about 30.
ZSM-5 is a shape-selective zeolite made by including a soluble organic template
in the mix of raw materials. Templates for this kind of synthesis include quarternary
ammonium salts. ZSM-5 enhances distillate yields in FCC units and catalytic dewax-
ing in hydroprocessing units, where due to its unique pore structure, it selectively
cracks waxy n-paraffins into lighter molecules.
The amorphous silica/alumina (ASA) catalysts used for hydrocracking are less
active, but they do a better job of converting straight-run VGO into diesel with
11.4 Petroleum Refining Catalysts 227
Mesopores
ASA Zeolite
Fig. 11.13 Schematic comparison of amorphous alumina silica (ASA) and zeolite structures
Finished FCC catalysts are powders with the consistency of sifted flour. They are
given their final form by spray drying in which a slurry of catalyst components is
converted into a dry powder by spraying into hot air. The method gives a consistent
particle size distribution. The particles are roughly spherical, with bulk densities of
0.85–0.95 g/cm3 and average diameters of 60–100 μm.
The catalysts include four major components—one or more zeolites (up to 50%), a
matrix based on non-crystalline alumina, a binder such as silica sol, and a clay-based
filler such as kaolin. The zeolite is ultra-stable (US) H-Y (faujasite) or very ultra-
stable (VUS) H-Y. The extra stability is provided by de-alumination with hydrother-
mal or chemical treatment. The H-Y zeolite is sometimes augmented with ZSM-5
to increase propylene yield. To provide thermal stability and optimize the relative
amounts of different active sites, a mixture of rare earths (RE), such as lanthanum-
rich mixture of rare earth (RE) oxides, is incorporated into the zeolite structure by
ion exchange. The RE mixture can contain up to 8 wt% ceria, up to 80% La2 O3 , up
to 15 wt% Nd2 O3 , with traces of other rare earths.
228 11 Cracking
Other components can be part of the core catalyst or introduced as external addi-
tives. The extra components can provide NOx and/or SOx reduction, Noble metal
combustion promoters enhance the conversion of CO to CO2 in the regenerator.
Fixed-bed catalysts are shaped into spheres, cylinders, hollow cylinders, lobed extru-
dates, pellets. Some look like small wagon wheels. Figure 11.14 shows some of these.
A cross-section of a lobed extrudate can look like a 3-leaf or 4-leaf clover without the
stem. Compared to cylindrical extrudates, shaped extrudates have a higher surface-
to-volume ratio, and the average distance from the outside of a particle to the center
is shorter. This increases activity by decreasing the average distances traveled by
molecules to reach active catalyst sites. To make extrudates, a paste of support mate-
rial forced through a die. The resulting spaghetti-like strands are dried and broken
into short pieces with a length/diameter ratio of 2–4; for main-bed hydrotreating cat-
alysts, diameters range from 1.3 to 4.8 mm. The particles are calcined, which hardens
them and removes additional water and volatile molecules such as ammonia.
Spherical catalysts are made by (a) spray-drying slurries of catalyst precursors, (b)
spraying liquid onto powders in a tilted rotating pan, or (c) dripping a silica-alumina
slurry into hot oil. Pellets are made by compressing powders in a dye. FCC catalysts
are made by spray drying.
Impregnation distributes active metals within the pores of a catalyst support. Like
sponges, calcined supports are especially porous. Far more than 99% of the surface
area is inside the pores. When the pores are exposed to aqueous solutions containing
active metals, capillary action pulls the aqueous phase into them. After drying, the
catalyst might be soaked in another solution to increase the loading of the same
(or a different) active metal. Catalysts can also be made by co-mulling active metal
oxides with the support. Co-mulling tends to cost less because it requires fewer steps.
It also produces materials with different activities—sometimes higher, sometimes
lower—than impregnation.
Eventually, refinery catalysts deactivate and must be replaced. The major causes
of deactivation are feed contaminants (trace metals, particulates, etc.) and catalyst
coking; the latter is discussed above in some detail. In fluid catalytic cracking (FCC),
continuous catalyst replacement (CCR) processes, and ebullated-bed hydrocracking,
aged catalysts are continuously removed and replaced with fresh. But in fixed-bed
units, catalyst replacement requires a shutdown. For a 40,000 barrels-per-day hydro-
cracker with an upgrade value of $15-20 per barrel, every day of down time for a
cost $600,000 to $800,000. Lost production during a 4-week catalyst changeout can
amount to $18 to $24 million.
11.5 Comparison of Catalytic and Thermal Cracking 229
Fig. 11.14 Catalyst loading scheme showing size/shape grading. Photo used with kind permission
from Haldor Topsoe Inc
Table 11.3 compares thermal cracking and catalytic cracking. The fundamental dif-
ferences are: thermal cracking does not use a catalyst. It operates at higher tempera-
ture; the pressure can be higher, as in steam cracking, or lower as in delayed coking.
The mechanism involves free radical chemistry. Catalytic cracking uses a catalyst
at lower temperature and pressure, and the mechanism involves ionic (carbocation)
reaction chemistry.
Table 11.4 compares the yields on similar topped crude feed by thermal and cat-
alytic cracking. Catalytic cracking produces more gasoline and olefins than paraf-
fins of the same carbon number. Thermal cracking produces more residual oil than
catalytic cracking. Catalytic cracking gives volume swell—the volume of liquid
products is greater than the volume of feed—but thermal cracking decreases liquid
230 11 Cracking
volume. Also, catalytic cracking produces coke that can deactivate the catalyst, but
thermal cracking of VGO does not.
From the crude distillation complex, straight-run gas oils and especially VGO can
be sent to hydrocracking units. Hydrocrackers produce LPG, light gasoline, heavy
naphtha, middle distillate fuels (jet and diesel), FCC feed, lube basestock, olefin
plant feed, and isobutane, which is an important feed component for alky plants.
Hydrocrackers also process cracked stocks, such as coker gas oils and FCC cycle
oils. Typically, if a refinery has both a hydrocracker and an FCC unit, straight-run
VGO goes to the FCC while cracked stocks go to the hydrocracker. In North America,
most hydrocrackers process cracked stocks.
Refineries which own recycle hydrocrackers and switch from maximum produc-
tion of middle distillate fuels in the winter to maximum production of naphtha in the
summer to meet heavier gasoline demands. Kerosene is a swing product, which can
be minimized or maximized by adjusting cut points. Hydrocracker middle distillate
are excellent blendstocks for diesel and jet. Hydrocracker naphtha has low octane,
but due to high naphthene content, it is a superb feed for catalytic reformers.
11.6 Hydrocracking 231
Table 11.4 Thermal versus Catalytic Cracking Yields on Similar Topped Crude Feed
Thermal cracking Catalytic cracking
wt% vol.% wt% vol.%
Fresh feed 100.0 100.0 100.0 100.0
Gas 6.6 4.3
Propane 2.1 3.7 1.3 2.2
Propylene 1.0 1.8 6.8 10.4
Isobutane 0.8 1.3 2.6 4.0
n-Butane 1.9 2.9 0.9 1.4
Butylene 1.8 2.6 6.5 10.4
C5 + gasoline 26.9 32.1 48.9 59.0
Light cycle oil 1.9 1.9 15.7 15.0
Decant oil 8.0 7.0
Residual oil 57.0 50.2
Coke 0 5.0
Total 100.0 96.5 100.0 109.9
C2 H5 SH + H2 → C2 H6 + H2 S (11.2)
Table 11.6 Comparison of heats of reaction (kJ/mol) calculated from bond energies with that heats
of formation for Reactions 11.1 and 11.2
Bonds broken Bonds formed Net bond energy Calc. from Hf
Reaction 11.1 C–C, H–H 2 C–H −42 −42.3
Reaction 11.2 C–S, H–H C–H, S–H −52 −58
1. C6 H14 + H2 → 2 C3 H8
2. C2 H5 SH + H2 → C2 H6 + H2 S
234 11 Cracking
C7H15 H2
CH3
+ C6H14
CH2CH3 Dealkylation CH2CH3
CH3 3 H2 Saturation
CH2CHCH2CH2CH3 H2 CH3
H2 Dealkylation
CH3 H2
C6H14 2 C 3H 8
+ C5H12
Paraffin
CH3 Hydrocracking
The chemical mechanism for catalytic hydrocracking is essentially the same as that
for FCC catalytic cracking. The processes themselves are considerably different
due to different catalysts, equipment, and operating conditions. An integral part of
hydrocracking is hydrotreating, which in fixed-bed units removes almost all oxygen,
sulfur, nitrogen and trace elements from the feed before it reaches the cracking
catalyst. Organic nitrogen poisons acidic cracking sites, so its removal is essential.
Along the way, hydrocracking also isomerizes n-paraffins and saturates olefins and
aromatics. With respect to equipment and process flow, fixed-bed hydrocrackers are
similar to fixed-bed hydrotreaters.
As shown in Fig. 11.16, hydrocrackers can have many different configurations; for
simplicity, pumps, heaters and heat exchangers are not shown. Most commercial units
have two reactors, but at least one unit (at Chevron El Segundo) has 6 reactors. Some
reactors have two beds. Other reactors have seven beds. Some units are once-through;
the unconverted oil goes off-plot. Others recycle unconverted oil to near-extinction.
Some have two flash drums, others have four. Some have amine treaters, others do
not. In some, only the gas is heated—hot gas is mixed with warm feed just before
the 1st reactor. In others, gas and oil are mixed before heating.
• Sketch 1 shows a once-through unit in which the oil flows in series through the
pretreat catalysts directly to the cracking catalysts.
• Sketch 2 shows two options for recycle of unconverted oil (UCO). The UCO can
go either to R1 (first reactor) or R2 (2nd reactor).
236 11 Cracking
Feed Makeup
Treat Gas Feed Treat Gas Makeup Gas
Gas
H2S Removal
H2S
Optional
PT Removal
1 PT 3
Recycled UCO
HC Seps
Seps HC
Strip
Strip
Products
Products
Frac
Frac
Products
Products
UCO
UCO
Strip, Frac
Products
Fig. 11.16 Hydrocracker Configurations (PT: pretreater, HC: hydrocracker, HP: high pressure, LP:
low pressure, UCO: unconverted oil). The makeup gas is hydrogen-rich with H2 ranging from 80
to 99.9%
• Sketch 3 shows a unit in which the pretreated oil is stripped or fractionated before
moving on to a hydrocracking reactor.
• Sketch 4 shows a unit with two independent recycle gas loops, one for PT and HC-1
and another for HC-2. This configuration has two major process advantages: in R3,
the pressure can be lower and the environment is nearly sweet—sweet because H2 S
and ammonia are stripped, and almost all of the nitrogen and sulfur are removed
in PT and HC-1. Temperatures required by uninhibited sweet hydrocracking are
lower. This leads to increased saturation of aromatics, which improves the quality
of middle distillate products. Some refiners still use noble-metal catalysts, due
to the higher activity and lower gas production over such catalysts. The Sketch
4 option seems like it would be more expensive, because it contains additional
equipment. But in one recent head-to-head comparison, due to the lower pressure
and less expensive metallurgy, the estimated installation cost of a Sketch 4 unit
was nearly the same as that for a Sketch 2 unit.
Figure 11.17 shows a two-reactor once-through hydrocracker with four catalyst
beds, along with some typical oil properties at various points in the process. The
liquid feed in this case is a typical straight-run VGO.
The makeup hydrogen can come from a steam-methane reformer (SMR), a cat-
alytic reformer, purified refinery purge gas, an olefins plant, or even an electrolysis
11.6 Hydrocracking 237
Fig. 11.17 Once-through single-stage (series flow) hydrocracking flow sketch with selected stream
properties
plant. The highest purity makeup (99.9 + % H2 ) comes from an SMR in which purifi-
cation is achieved with a pressure swing adsorption (PSA) unit. Makeup gas from an
SMR equipped with Benfield purification can contain 4 vol.% methane. Olefin plant
hydrogen has up to 5 vol.% methane, and electrolysis hydrogen can contain trouble-
some traces of HCl. Purified refinery purge gas can have purities ranging from 90 to
95% H2 for a membrane unit to 99.9 + % H2 for a PSA. FCC and thermal cracking
units produce a lot of low-purity hydrogen containing olefins; it is uneconomical to
purify such gases.
Hydrotreating catalysts are described above. They remove metals, organic sulfur,
and organic nitrogen. Sulfur doesn’t harm the hydrocracking catalysts in R2, but
nitrogen inhibits acid-induced cracking.
Hydrocracking catalysts are bifunctional, containing metals for saturation and
a solid acid for cracking. The active metals can be either palladium, which is an
expensive noble metal, or a base-metal sulfide (MoS2 or WS2 ) promoted by NiS.
The acid is either an amorphous aluminosilicate or a crystalline zeolite. These were
discussed in the refining catalyst section.
A mixture of liquid feed and hydrogen-rich gas enter the first reactor at 550 °F
(start-of-run) to 750 °F (end-of-run); these temperatures correspond to 290 and
400 °C, respectively. Temperatures increase by as much as 100 °F (56 °C) but in
no case are they allowed to exceed the metallurgical limit (usually 825 °F, 440 °C).
Between beds, relatively cold quench gas is added to reduce the temperature prior to
the subsequent bed. The effluent from the last first-stage reactor is sent through a heat
exchanger train, to a series of two to four separators (flash drums). Before the cold
238 11 Cracking
high-pressure separator (CHPS), wash water is added to dissolve NH4 SH and NH4 Cl,
which otherwise would foul tubes in the reactor effluent air cooler (REAC). The
resulting sour water is drained from the boot of the CHPS. In the cold high-pressure
separator, hydrogen laden with H2 S, methane, and other light gases is recovered
overhead and recycled. Sometimes, an amine unit is used to remove H2 S from the
recycle gas. The high-pressure separator bottom stream goes to low-pressure sepa-
rator(s), which remove residual gases before the liquid product proceeds to stripping
and fractionation. Some of the recycle gas is purged to maintain purity. The purge
stream, which contains 80–85% hydrogen, can go to hydrotreaters. As mentioned
above, purge gases can be purified with a membrane unit which boosts the purity of
diffused (low-pressure) gas to >95%. Purification with a PSA provides makeup with
99.9 + % H2 .
From the separators, the liquids go to a fractionator, which cuts the full-range
product into individual streams—light naphtha, heavy naphtha, middle distillates,
and unconverted oil (UCO).
There can be one middle distillate draw (diesel) or two (kerosene and diesel). The
fractionator can be operated to give sales-quality diesel. Kerosene quality is highly
feed dependent. It often meets sales specifications, but toward the end of a catalyst
cycle, high aromatics might keep it from meeting smoke-point specifications.
As mentioned, the UCO can go off-plot to an FCC unit or olefins plant. It can
also be used as lube base stock. The UCO from hydrocracking is a premium product.
Compared to the VGO from the vacuum distillation column, it has higher isoparaffin
concentration, high viscosity index, less sulfur, less nitrogen, lower aromatics, and
higher hydrogen content. When recycled to the reactors for additional conversion,
overall conversions can exceed 98 wt%.
The hydrocracker is a pivot point in the refinery—the “process in between.” It
takes straight-run VGO from the crude distillation complex, but it also takes feed
from other conversion units. It generates finished products—light naphtha and middle
distillates—but it also provides feeds to other units.
Feeds Straight-run VGO, coker gas oil, FCC cycle and slurry oils, DAO,
extract.
Intermediates Isobutane => alkylation
Light naphtha => gasoline blending
Heavy naphtha => catalytic reforming
Middle distillates => jet, diesel
Unconverted oil (UCO) => recycle, FCC, olefins plant, lube bases-
tock
As shown in Table 11.7, a fixed-bed recycle hydrocracking unit can have signifi-
cant product flexibility, producing either large amounts of C4 -plus naphtha or large
amounts of middle distillates. In petroleum refining, this kind of process flexibility
is unique.
Catalyst cycles last from 1 to 4 years, typically for two years. Units run to achieve
specified targets, such as “this much” conversion or “that much” production of FCC
11.6 Hydrocracking 239
or olefin-plant feed. As a catalyst cycle progresses, it’s necessary to raise the average
temperature about 1–3 °F per month to compensate for loss of catalyst activity.
Higher hydrogen partial pressure improves almost every aspect of hydrocracker
operation. The maximum operating pressure of an existing unit is fixed, but hydrogen
partial pressure can be increased by purging recycle gas or improving makeup gas
purity.
Conversion is a function of residence time, i.e., feed rate. In an existing unit,
increasing the feed rate decreases conversion. To compensate, it’s necessary to
increase temperature.
An increase in feed nitrogen content might decrease conversion, because organic
nitrogen inhibits catalyst activity. However, pretreat reactors are operated to maintain
constant nitrogen in the feed to the hydrocracking catalyst, typically 10–30 ppmw,
so the impact of feed nitrogen on the cracking catalyst is dampened.
If a unit is not designed for high levels of feed sulfur, corrosion of equipment
could impair unit performance. Feed sulfur affects the pretreat catalyst more than it
does the hydrocracking catalyst. For both catalysts, removing hydrogen sulfide from
the recycle gas with an amine treater diminishes the impact of feed sulfur content.
The buildup of heavy polynuclear aromatics from recycled fractionator bottoms
can cause fouling of heat exchanger. Reducing feed end point or removing 2% of the
UCO in a drag stream may be necessary.
Heavy feeds, such as residual fuel oils and reduced crudes contain high concentrations
of asphaltenes and ash, and they contain more sulfur, nitrogen, and metal-containing
components (such as metalloporphyrins) than gas oils. Typically, catalytic residue
upgrading processes are applied to atmospheric residues (AR). The type of catalysts
and operation conditions in AR hydrocracking are different than those used for
gas oils. Vacuum residues (VR) have low hydrogen/carbon ratio and high metal
content, which deactivates catalysts rapidly. They typically are processed in non-
catalytic processes, such as solvent extraction, delayed coking, Flexicoking, and
thermal hydrocracking. (However, certain thermal hydrocracking additives can have
catalytic activity.)
In contrast to fixed-bed hydrocrackers, ebullated bed (e-bed) units can process
large amounts of atmospheric and vacuum residues (AR and VR) with high metals,
sulfur, nitrogen, asphaltenes and solid contents. They employ catalysts with both
hydrotreating and hydrocracking activity; in such units, it is impossible to segregate
catalysts, so the catalyst accomplishes both hydrotreating and hydrocracking. As dis-
cussed, hydrotreating removes sulfur and nitrogen and hydrogenates aromatic rings.
Hydrocracking entails catalytic breaking of C–C bonds. E-bed processes convert
residue-containing feeds into distillates and upgraded bottoms for FCC and other
conversion units.
11.6 Hydrocracking 241
Process-wise, catalyst life does not limit these units, because fresh catalyst is
continually added as spent catalyst is removed. Economics determine whether or not
the benefits of residue conversion offset the cost of catalyst replacement.
In ebullated bed reactors, hydrogen-rich recycle gas is bubbled up through a mix-
ture of oil and catalyst particles. This provides three-phase turbulent mixing, which
is needed to ensure a uniform temperature distribution. The process can tolerate sig-
nificant differences in feed quality, because in addition to manipulating temperature,
operators can change catalyst addition rates. Catalyst consumption is determined by
the concentrations of trace metals—particularly Fe, Ni and V—in the feed.
Figure 11.18 shows an H-Oil reactor that uses ebullated-bed hydrocracking tech-
nology to process heavy feedstock residues such as vacuum gasoils (VGO), deas-
phalted oils (DAO), and Coal derived oils. A fresh catalyst is continuously added and
the spent catalyst withdrawn to control the level of catalyst activity in the reactor,
enabling constant yields and product quality over time.
At the top of the reactor, catalyst is disengaged from the process fluids, which are
separated in downstream flash drums. Most of the catalyst is returned to the reactor.
Some is withdrawn and replaced with fresh catalyst. When compared to fixed-bed
processes e-bed technology offers the following advantages:
• The ability to achieve more than 70 wt% conversion of atmospheric residue.
242 11 Cracking
• Ample free space between catalyst particles, which allows entrained solids to pass
through the reactor without accumulation, plugging, or build-up of pressure drop.
• Better liquid-product quality than delayed coking.
Disadvantages versus fixed-bed processes include high catalyst attrition, which
leads to high rates of catalyst consumption; higher installation costs due to larger
reactor volume and higher operating temperatures; and sediment formation. Recent
improvements include second-generation catalysts with lower attrition; catalyst reju-
venation, which allows the reuse of spent catalysts; improved reactor design leading
to higher single-train feed rates; and two-reactor layouts with inter-stage separation.
Another version of fluidized (or extended) bed hydrocracking is LC-fining devel-
oped by Lummus (now CB&I), shown in Fig. 11.19.
Off Gases
Sulfur
Etc.
Stage 2
Hot HP Reactor
Vacuum Resid Separator
Stage 1 Gas
Reactor Processing
Light
Ends
Additive
Cold Naphtha
Separator
Vacuum Middle
Flash Distillates
Hydrogen Residue
UCO
Fig. 11.20 Two-stage slurry-phase hydrocracking process flow. Based on drawings supplied by
KBR Technology. Used with kind permission from KBR, Inc. [23]
Inside the reactor, the liquid/additive mixture behaves as a single phase due to the
small size of the additive particles. The additives prevent bulk coking by providing
highly dispersed nucleation sites for “micro coking.” The additive isn’t recovered.
Instead, it ends up in a pitch fraction, which comprises <5 wt% of the feed.
Slurry-phase hydrocracking has several advantages:
• It can achieve >95 wt% conversion of vacuum residue and high conversion of coal.
• In two-stage designs, which incorporate fixed-bed hydrotreating and hydrocrack-
ing, product quality is excellent.
• Feeds can include vacuum residue, FCC slurry oil, and even coal.
• The additive is low-cost and disposable.
• For a given volume of residue feed, total slurry phase reactor volume is lower than
the reactor volume of e-bed processes.
The main disadvantage of the KBR slurry-phase hydrocracking is the unconverted
pitch. The pitch quality is so low that it is exceptionally difficult to dispose.
A better process is LC-MAX, offered by CB&I. LC-MAX is a combination of
LC FINING and solvent deasphalting. The conversion to fuels is somewhat lower,
but all products, including the asphalt, can be sold conventionally.
References
1. Greensfelder BS, Voge HH, Good GM (1949) Catalytic and thermal cracking of pure hydro-
carbons. Ind Eng Chem 41(11):2573–2584
2. Zhu G, Xie C, Li Z, Wang X (2017) Catalytic processes for light olefin production (Chapter 36).
In: Hsu CS, Robinson PR (eds) Springer handbook of petroleum technology. Springer, New
York
244 11 Cracking
3. https://www.quora.com/What-does-a-steam-cracker-look-like-and-what-are-its-essential-
components. Accessed 6 Aug 2014
4. Hsu CS, Robinson PR (2017) Gasoline production and blending (Chapter 17). In: Hsu CS,
Robinson PR (eds) Springer handbook of petroleum technology. Springer, New York
5. Speight J (2017) Fluid-bed catalytic cracking (Chapter 19). In: Hsu CS, Robinson PR (eds)
Springer handbook of petroleum technology. Springer, Heidelberg
6. Avidan A, Edwards M, Owen H (1980) Experiments performed at constant coke production.
Oil Gas J 88(1):52
7. von Ballmoos R, Harris DH, Magee JS (1995) Catalytic cracking (Section 3.10). Encyclopedia
of Catalysis. Wiley, London, pp 1955–1985
8. Van Antwerpen FJ (1944) Thermofor catalytic cracking. Ind Eng Chem 36(8):694–698
9. AIChE (2009) Chemical engineers and energy: In: Chemical engineers in action—innovation
10. Palmas P (2009) Traces of the history of RFCC and provides guidelines for choosing the appro-
priate regenerator style: https://www.uop.com/?document=uop-25-years-of-rfcc-innovation-
tech-paper&download=1. A reprint from hydrocarbon engineering
11. Park JI, Mochida M, Marah AMJ, Al-Mutairi A (2017) Modern approaches to hydrotreat-
ing catalysis (Chapter 21). In: Hsu CS, Robinson PR (eds) Springer handbook of petroleum
technology. Springer, New York
12. Resid Upgrading, By Honey UOP: https://www.uop.com/processing-solutions/refining/
residue-upgrading/#resid-fcc. Accessed 31 Aug 2016
13. Honeywell UOP, “Catalyst Cooler,” https://www.uop.com/equipment/fcc/catalyst-cooler/.
Accessed 1 Nov 2018
14. Prime-G +: The benchmark technology for ultra-low sulfur gasoline by Axens: https://zh.
scribd.com/doc/209103920/Prime-G. Accessed 20 Sep 2106
15. Laan JV ConocoPhillips S Zorb Gasoline Sulfur Removal Technology: http://www.icheh.com/
Files/Posts/Portal1/S-Zorb.pdf. Accessed 20 Sep 2016
16. Schroder M (2010) Functional metal organic frameworks: gas storage. Sep Catal 293:175–205
17. Olah GA (1994) My search for carbocations and their role in chemistry. In: Nobel Lecture
18. Robinson PR, Dolbear GE (2017) Hydrocracking (Chapter 22). In: Hsu CS, Robinson PR (eds)
Springer handbook of petroleum technology. Springer, New York
19. Bergius Process, Project Gutenberg, http://self.gutenberg.org/articles/Bergius_process.
Retrieved 4 Aug 2015
20. Lapinas AT, Klein MT, Gates BC, Macris A, Lyons JE (1991) Catalytic hydrogenation and
hydrocracking of fluorene: reaction pathways, kinetics, and mechanisms. Ind Eng Chem Res
30(1):42–50
21. https://archive.epa.gov/emergencies/docs/chem/web/pdf/tosco.pdf. Retrieved 9 Oct 2018
22. Chevron Technology Marketing: LC Fining: http://www.chevrontechnologymarketing.com/
CLGtech/lc_finishing.aspx Retrieved 4 Aug 2015
23. Motaghi M, Ulrich B, Subramanian A (2011) Slurry-phase hydrocracking: possible solution
to refining margins. Hydrocarbon Proces 90(2):37
Chapter 12
Coking and Visbreaking
12.1 Coking
Delayed coking is a cyclic process that employs more than two coke drums—typ-
ically at least four, as a semi-continuous or semi-batch process. The drums operate
on staggered 18–24 h cycles, during which a drum is full for about 12–18 h. Each
cycle includes preheating the drum, filling it with hot oil, allowing coke and liquid
products to form, cooling the drum, and decoking. A typical coking unit with two
coke drums is shown in Fig. 12.1.
The feedstock of the coking unit (coker) is typically residue from the vacuum
distillation unit. It goes to the coke furnace to be heated to about 900–970 °F
(487–520 °C) at 90 psi pressure. At this stage, steam is injected to minimize cracking
of heavy liquid into smaller molecules as gas and liquid products. The products of
coker furnace continue to be pumped into a large drum (soaker) . Thermal cracking
12.1 Coking 247
begins immediately, generating coke and cracked products. Coke accumulates in the
drum at 845–900 °F while the vapors rise to the top then go to a product fractionator.
Meanwhile, hot feed keeps flowing into the drum until it is filled with solid coke.
When gas formation stops, the soaking step is complete. At this point, the top and
bottom heads of the drum are removed. A rotating cutting tool uses multiple high-
pressure jets of water (500–3500 psi) to drill a hole through the center of the coke
from top to bottom as a pilot nozzle for decoking. In addition to cutting the hole,
the water also cools the coke, forming steam as it does so. Next, the high-pressure
water is switched from vertical pilot nozzle to horizontal reamer. The cutter (pilot
nozzle and reamer) is then raised, step by step, (or moved up and down vertically)
cutting the coke into lumps, which fall out the bottom of the drum. The wet coke
goes through dewatering process to final dry coke product.
Light products include gases, coker naphtha, light coker gas oil (LCGO), and
heavy coker gas oil (HCGO). All of these require further processing due to their
high content of sulfur, nitrogen, and olefins, which makes them unstable and poorly
suited for direct blending into finished products. The gases are treated with amine
to remove hydrogen sulfide. The coker naphtha and LCGO are hydrotreated. The
HCGO can go either to an FCC unit or a hydrocracker.
Coke accounts for up to one-third of the product. The raw coke from the coker is
referred to as green (unprocessed) coke. It can be upgraded to fuel grade (relatively
high in sulfur and metals) and anode grade (low in sulfur and metals). Desirable forms
include sponge coke and needle coke. Shot coke is dangerous and undesirable.
Fluid coking, also called continuous coking, is a moving-bed process for which the
operating temperature is higher than the temperatures used for delayed coking. The
process burns only enough of the coke to satisfy the heat requirement of the reactor
and feed preheat. It typically consumes 20–25% of the coke produced in the reactor.
The rest of the coke is withdrawn from the burner vessel, which can be used as
furnace fuel in the refinery. The first commercial fluid coker was built at the Billings,
Montana refinery of the Carter Oil Company.
In continuous coking, hot recycled coke particles are combined with liquid feed
in a radial reactor at about 50 psig (446 kPa) as shown in Fig. 12.2. The feed, such
as vacuum residue is preheated to 500–700 °F before entering the coker, which is
operated at 900–1050 °F at 10 psi pressure with recycled hot coke from a burner at
1100–1200 °F.
Vapors are taken from the top of the reactor, quenched to stop any further reaction
and fractionated. Metals and Conradson carbon residue in the feed are rejected with
the coke, but the liquid products still contain sulfur and nitrogen. The VGO goes to a
hydrocracker or an FCC feed pretreater. The hydrotreated naphtha goes to blending or
a catalytic reformer, and the hydrotreated middle distillates go to blending. Operating
conditions are selected for maximum yield of distillate products. Less coke is made
248 12 Coking and Visbreaking
than in delayed coking. The coke goes to a burner, where coke is partially burned
to provide sufficient heat for cracking: the heat is transferred when the hot coke
particles return to the coker. The remaining unburned coke (up to 75%) cannot be
sold as product, but is purged out of the coker as fuel for other parts of refinery,
such as furnaces. The yield of distillates from coking can be improved by reducing
residence time. However, fluid cokers can be designed and operated to produce CO
and H2 from the coke. Installation costs for fluid coking are considerably higher than
for delayed coking, but feeds can be heavier, heat losses are lower, and conversion is
higher. Unfortunately, much of the incremental conversion is to C3 - material. Fluid
coking makes more fuel gas than delayed coking.
products and coke. The coke circulates from the reactor to the heater via the cold
coke transfer line. In the heater, the coke is heated by the gasifier gas (a low BTU
mixture of CO/H2 /CO2 ) and circulated back to the reactor via the hot coke transfer
line to supply the heat that sustains the thermal cracking reaction. The excess coke in
the heater is transferred to the gasifier, where it reacts with air and steam to produce
a gas. The gasifier products, consisting of a gas and coke mixture, return to the heater
to heat up the coke. The gas exits the heater overhead and goes to steam generation
and to the integrated FLEXSORB™ desulfurization process to reduce H2 S in the
flue gas. The low BTU gas is typically fed to a CO boiler for heat recovery but can
also be used in modified furnaces/boilers; atmospheric or vacuum distillation tower
furnaces; reboilers; waste heat boilers; power plants and steel mills; or as hydrogen
plant fuel, which can significantly reduce or eliminate purchases of expensive natu-
ral gas. The small amount of residual coke produced can be sold as boiler fuel for
generating electricity and steam or as burner fuel for cement plants.
12.2 Visbreaking
Visbreaking is a mild form of thermal cracking that achieves about 15% conversion
of atmospheric residue into gas oils and naphtha. At the same time, a low-viscosity
residual fuel is produced. The primary purpose of visbreaking is not for conversion,
but to reduce the viscosity and pour point of fuel oil, such as No. 6 fuel oil, to
250 12 Coking and Visbreaking
acceptable levels for specifications. The reaction is quenched before the residual fuel
can form coke. The two main types of visbreaking are “short-contact” and “soaker.”
In short-contact coil visbreaking, cracking occurs in furnace tubes (coils) and the
product quenched by a stream of cold oil. As shown in Fig. 12.4, the feed is heated
to about 900 °F (480 °C) in furnace tubes (coils) and sent to a soaking zone reactor
at 140–300 psig (1067–2170 kPa). Steam (~1 wt%) is injected along with the feed
to generate turbulence, control the liquid residence time and prevent coking along
the tube wall. The elevated pressure allows cracking to occur while restricting coke
formation. To avoid over-cracking, the residence time in the soaking zone is short-
—several minutes (typically 1–3 min) compared to several hours in a delayed coker.
The hot oil is quenched by heat exchange with feed or cold gas oil to inhibit further
cracking and sent to a vacuum fractionator for product separation.
Soaker visbreaking, shown in Fig. 12.5, keeps the hot oil at relatively lower temper-
ature, about 800 °F (430 °C), and at 50–200 psi in a soaker drum after the furnace for
a longer time (15–25 min) to allow more cracking to occur before quenching, This
increases the yield of middle distillates. Low-viscosity visbreaker gas oil can be sent
to an FCC unit or hydrocracker for further processing, or used as heavy fuel oil.
Table 12.1 compares the product yield for catalytic and thermal cracking processes.
The catalytic cracking is represented by FCC and hydrocracking, and thermal crack-
ing by coking, visbreaking and steam cracking. Thermal cracking to produce gaso-
line has been replaced by catalytic cracking. However, steam cracking has become a
primary method for producing ethylene and propylene for the polymer production,
while catalytic cracking is the key method for producing fuels from heavy distillates.
cracked or
visbroken residue
Table 12.1 Comparison of product yields for catalytic and thermal cracking processes [6]
Process Type Product characteristics
FCC Catalytic cracking in the C1 and C2 : low to moderate
absence of external H2 Iso/normal paraffin ratio: high
C3 and C4 olefin yields: can be significant
H2 production: moderate
Aromatics: higher than feed
Coke formation: high
Alkyl aromatics: β-scission next to the ring
Catalytic Catalytic cracking in the C1 and C2 : low
hydrocracking presence of external H2 Iso/normal paraffin ratio: high
Olefins: removed by pretreating
Aromatics: significantly lower than feed
Coke formation: minimal
Alkyl aromatics: β-scission next to the ring
Slurry-phase Thermal cracking in the C1 and C2 : high
hydrocracking presence of external H2 Iso/normal paraffin ratio: similar to feed
Olefin production: low
Aromatics: depends on feed
Coke formation: low
Alkyl aromatics: scission within the side
chain
Coking/visbreaking Thermal cracking C1 and C2 : high
Iso/normal paraffin ratio: similar to feed
Olefin production: high
Aromatics: moderate
Alkyl aromatics: scission within the side
chain
Steam cracking Thermal cracking in the Olefin production: high
presence of steam H2 production: high
Aromatics production: high
Coke formation: low
References
The use of petroleum heavy fractions for lubrication started in early civilization. For
example, Egyptians use pitch for lubricating chariot wheels more than 4000 years
ago. Today, there are many uses for lubricants and specialty oils, with some examples
listed below:
Automotive Engine oils, automatic transmission fluids (ATF’s), gear oils.
Industrial Machine oils, greases, electrical oils, gas turbine oils.
Medicinal Food grade oils for ingestion, lining of food containers, baby oils.
Specialty Food grade waxes, waxes for candles, fire logs, cardboard.
Normal paraffins have the higher VI, but also have the highest pour point, discussed
below. Hence, they must be removed by solvent dewaxing or catalytic dewaxing
processes.
Pour point is the temperature at which the fluid ceases to pour and is nearly a
solid. Low pour point ensures ease of starting and proper start-up lubrication on cold
days. Viscosity pour point is measured by lowering temperature until flow stops. Wax
pour point occurs abruptly as paraffinic wax crystals precipitate from solution and
the oil solidifies.
Cloud point is the temperature at which wax crystals first appear.
Boiling range is a factor along with viscosity for the selection of cut points for
lubes.
Volatility is a measure of oil loss due to evaporation. The volatility is generally
lower for higher viscosity and higher VI base stocks.
ASTM D5800 (Noack method) [1] is the commonly accepted method for mea-
suring volatility of automotive lubricants. The test simulates the evaporation loss in
operating internal combustion engine for high-temperature services.
Simulated distillation (SimDist) provides a convenient means of measuring the
front end of boiling point curves for volatility measurements. ASTM D2887, ASTM
D6352 and ASTM D7213 are commonly used [1].
Flash Point is a measure of the temperature at which there is sufficient vapor in
the oil to ignite, which also reflects the front of the boiling point curve. Flash point
is a required safety specification for many base stocks. ASTM D92 can be applied
for the measurements.
Saturates, naphthenes and aromatics are the content of these molecular types
present in the base stock. They are determined by chromatography and mass spec-
trometry at the molecular level [2].
Oxidative resistance is the resistance to rapid oxidation at high temperature, espe-
cially when oil comes in contact with the piston head. Oxidation causes the forma-
tion of coke and varnish-like asphaltic materials from paraffinic base oils and sludge
from naphthenic base oils. Phenolic additives and zinc dithiophosphates (ZDDP’s)
are added to suppress oxidation and its effects.
Acidity causes corrosion of bearing metals. Acids are formed from oxidation of
lube oil hydrocarbons and introduced into the crankcase by piston blow-by of engine
combustion by-products. Paraffinic base oils exhibit lower acidities than naphthenic
base oils due to excellent thermal and oxidation stabilities.
Color is commonly measured by ASTM D1500 [1]. Most premium base stocks
have a D1500 reading less than 0.5 when they are fresh. An increase in the D1500
reading is an indication of formation of oxidation products. However, many manu-
facturers have their own proprietary methods for measuring oxidation and storage
stability of base oils.
Conradson carbon (CCR) or Micro-Carbon Residue (MCR) is a measure of the
ash left after flame burning.
256 13 Lubricant Processes and Synthetic Lubricants
Lubes made from virgin distillates and solvent extracts are usually referred to neu-
trals, such as 100 N (or S100 N), 150 N, 600 N, etc. Bright stock comes from Deas-
phalted Oil (DAO) from vacuum resid. An example is BS150. The grades shown in
the finished oils are expressed as SAE 5, 10, 30, or ISO 22, 32, etc.
The Society of Automotive Engineers (SAE) developed a scale for both engine
oils and transmission oils. The multi-grade notation as “xWy” is commonly used
for the motor oils on shelves today. “x” refers to viscosity when cold where “W’
stands for winder grade, which rates the oil’s flow at 0 °F (−17.8 °C). The lower
the number, the less the oil thickens in the cold. A low number before “W”, such
as 0W20 or 5W30, mean that engines start more easily and are better protected in
cold weather. “y”, on the other hand, refers to viscosity measured at 212 °F (100 °C),
which rates the oil’s resistance to thinning at high temperatures. The higher the
number, the more readily the oil thins. Hence 20W-30 will thin more than 20W-40
at higher temperatures. The difference between the x and y values reflects viscosity
index (VI). Low differences indicate higher VI. Monograde designations, such as
SAE 20, 30, or 40, are no longer used.
The American Petroleum Institute (API) organizes lube base stocks into five
categories, listed in Table 13.1. In general, Groups I–III base stocks are from crude
oils and are made through solvent refining or hydroprocessing (hydrocracking or
hydroisomerization), while Groups IV and V are synthetic.
Paraffinic crudes, such as West Texas and Arab Light, have higher wax content
and require n-paraffin removal, via isomerization or dewaxing, to remove the wax.
Naphthenic crudes, such as Venezuelan and many California crudes, have low wax.
Figure 13.1 shows how lubricant feedstocks are derived from crude oil fractions
[3]. The reduced or topped crude from the atmospheric distillation tower is sent to a
vacuum distillation tower to fractionate the feed into light vacuum gas oil (LVGO)
which can be used as distillate fuels after further upgrading, heavy vacuum gas oil
(HVGO) for lube processing, and vacuum resid (residue) which can also be used as
lube feedstock after deasphalting.
The lube plant can also have its own vacuum tower, taking reduced crude from the
refinery, shown in Fig. 13.2. It produces several lube feedstock fractions for different
grades of lube oils—light, medium and heavy. Bright stock comes from vacuum
resid.
258 13 Lubricant Processes and Synthetic Lubricants
Among all the crude oils available, the first step is to identify economically attractive
crudes through characterization and assays that include lube yields and product qual-
ity. The lube assays include atmospheric distillation, vacuum distillation, extraction
and dewaxing. Computer modeling using assay information is employed to predict
process response for desired lube products and to investigate process variables and
operating optimization for distillation, extraction, and dewaxing. An important goal
is to assess manufacturing flexibility. Even the best models have limitations, so a
plant test is performed, including blending into a formulated lube for further test-
ing. Periodic re-evaluation is needed to assess lube quality variation over time due
to feed changes and process changes caused by equipment fouling. During ongo-
ing economic evaluation, lube yields and product qualities are evaluated. Important
parameters include viscosity, sulfur, density, refractive index, oxidative stability, etc.
The viscosity index of hydrocarbon types is listed in Table 13.2. Normal paraffins
have the highest viscosity index, but those boiling in the lube range form wax deposits,
especially in cold weather. Hence, dewaxing is an important process to remove normal
paraffins while retaining isoparaffins and mononaphthenes that also have high VI.
Aromatics have the lowest VI. They are removed by extraction from lube base stock,
along with polynaphthenes.
The viscosity index can be calculated using the following formula:
where U is the oil’s kinematic viscosity at 40 °C (104 °F), and L and H are values
based on the oil’s kinematic viscosity at 100 °C (212 °F). L and H are the values of
viscosity at 40 °C for oils of VI = 0 and 100 respectively, having the same viscosity
at 100 °C as the test oil. L and H values can be found in ASTM D2270 [2]. If the
kinematic viscosity of the oil is less than or equal to 70 mm2 /s (cSt), VI can be
obtained by linear interpolation. If the kinematic viscosity is above 70 mm2 /s (cSt),
L and H are calculated by the equation below:
where:
L Kinematic viscosity at 40 °C of an oil of VI = 0 having the same kinematic
viscosity at 100 °C as the oil whose VI is to be calculated,
Y Kinematic viscosity at 100 °C of the oil whose VI is to be calculated,
H Kinematic viscosity at 40 °C of an oil of VI = 100 having the same kinematic
viscosity at 100 °C as the oil whose VI is to be calculated.
There are several processes that can be used for viscosity index improvements,
including:
• Hydrogenation via hydrotreating to convert polyaromatic hydrocarbons to poly-
naphthenic ring compounds.
• Hydrocracking of VGO to increase paraffin concentration and viscosity index.
• Higher severity hydrocracking to convert polyaromatics into mononaphthenes and
isoparaffins.
• Hydroisomerization (Isodewaxing) to catalytically convert C20 + normal paraffins
to C20 + isoparaffins with minimum cracking.
• In both hydrogenation and hydroisomerization, dehydrogenation of naphthenes to
aromatics should be avoid or minimized.
• Separation of aromatic hydrocarbons from paraffins and naphthenes through sol-
vent extraction.
Figure 13.3 shows a flow diagram for a conventional solvent refining lube base stock
production plant. Solvent refining processes include solvent deasphalting, solvent
extraction, and deoiling/dewaxing processes. Vacuum distillation of atmospheric
residue yields vacuum residue, which undergoes propane deasphalting to remove
asphalt. Solvent deasphalting reduces coke and sludge formation in the bright stock.
The deasphalted oil goes to a solvent extraction tower, which removes most aromat-
ics. Lighter lube fractions from different levels of the vacuum tower also undergo
solvent extraction. The low-aromatics raffinates undergo solvent dewaxing. Solvent
extraction improves viscosity index by removing aromatics. Solvent dewaxing and
lowers cloud and pour points of the base stocks. Hydrotreating (hydrofinishing) or
clay treating improves color and oxidation stability and lowers organic acidity. The
wax is a useful and marketable by-product. It is refined through wax deoiling, wax
hydrotreating, or hydrofinishing.
260 13 Lubricant Processes and Synthetic Lubricants
Lubricant feedstocks come from heavy distillation fractions, including heavy gas
oil, atmospheric resid (topped crude), vacuum gas oils and vacuum resid, which may
contain significant amounts of asphaltene. Deasphalting becomes the first step of
lubricant oil manufacturing. Also, solvent deasphalting takes advantage of the fact
that undesirable aromatic compounds are insoluble in paraffins.
Propane or n-pentane are commonly used in industry to precipitate asphaltenes
from residual oils. The deasphalted oil (DAO) is sent to hydrotreaters, FCC units,
hydrocrackers, or fuel-oil blending. In FCC units, DAO is easier to process than
the corresponding straight-run residue. This is because, by definition, the asphaltene
content of DAO is very low. The asphaltenes in straight-run residue easily form coke
and often contain catalyst poisons such as nickel and vanadium. In hydrocrackers,
DAO can be harder to process than straight-run VGO and FCC cycle oils, because
although DAO no longer contains asphaltenes, it is rich in resins and hence still has
a very high endpoint and high CCR.
The earliest commercial applications of solvent deasphalting used propane as
the solvent to extract high-quality lubricating oil bright stock from vacuum residue,
shown in Fig. 13.4. In propane deasphalting, the residual oil and propane are pumped
to an extraction tower at 150–250 °F (65–120 °C) and 350–600 psig (2514–4240 kPa).
Separation occurs in a tower, which may include a rotating disc contactor. The raf-
finate flowing off at the top of extraction tower contains DAO, which goes to a
high-pressure flash tower to recover propane, which is recycled. The DAO then goes
to a stripper to get rid of other dissolved impurities. The final DAO product can
become bright stock in a lube plant, or sent as a feed for resid upgrading. With
substantially lower levels of metals and carbon contaminants, DAO is far easier to
13.8 Solvent Deasphalting 261
process than conventional resid. The bottoms of the extractor go to a high pressure
then a low-pressure flash tower to recover additional dissolved propane for recycle.
Asphalt is obtained from an asphalt stripper. In this process, DAO is considered as
an extract while asphalt is a raffinate.
An advanced version of solvent deasphalting is “residuum oil supercritical extrac-
tion,” or ROSETM . The ROSE Process was developed by the Kerr-McGee Corpo-
ration and now is offered for license by KBR. In this process, residue and solvent
are mixed and heated to above the critical temperature of the solvent. Liquid yields
are higher under supercritical conditions, because the lighter part of the oil becomes
more soluble. In addition to giving higher yields, the process is more energy efficient
and has lower operating costs due to improved solvent recovery. The ROSE pro-
cess can employ three different solvents, the choice of which depends upon process
objectives:
Propane Preparation of lube base stocks
Butane Asphalt production
Pentane Maximum recovery of liquid.
The ROSE process can make poor quality asphalt or provide feeds for FCC,
hydrocracking, or coking units. It can also be a unit in a lube complex. Unfortunately,
due to its high residue content, ROSE DAO is especially difficult to process in
conventional hydrotreaters and hydrocrackers.
262 13 Lubricant Processes and Synthetic Lubricants
Figure 13.5 shows typical molecules processed in the lube processing plants. Extrac-
tion eliminates the undesirable molecules and dewaxing eliminates wax, leaving
molecules between the two sloped lines.
Lube hydroprocessing includes three steps. First, hydrotreating removes
heteroatom-containing compounds and saturates aromatics and poly-aromatics. The
removal of nitrogen is required prior to the second step: hydrocracking. During
hydrocracking, additional aromatics are hydrogenated, mono-naphthenes are con-
verted into isoparaffins by ring opening, and multi-ring naphthenes are converted into
naphthenes with fewer rings, also by ring opening. With conventional hydrocracking
catalysts, some n-paraffins are transformed into i-paraffins. With isodewaxing cata-
lysts, large amounts of n-paraffins are isomerized. The third step is hydrofinishing,
which removes the last traces of sulfur and olefins, thereby stabilizing the finished
base stock.
Solvent extraction is used to remove aromatics and other impurities from the paraf-
finic and naphthenic portion of lube base stocks and grease stocks. This improves
viscosity index, color and oxidation resistance of the base stock, and reduces carbon
and sludge formation. As shown in Fig. 13.6, the feedstock is dried, then contacted
with the solvent in a counter-current or rotating disk extraction unit. The undesirable
polyaromatics are accumulated in solvent-rich extract phase. The desirable paraffinic
and naphthenic components are accumulated in light raffinate phase, which can be
hydrotreated prior to dewaxing. The solvent is separated from the product stream by
heating, evaporation, or fractionation. Remaining traces of solvent are removed from
the raffinate by steam stripping or flashing. Electrostatic precipitators can enhance
the separation of inorganic compounds. The solvent is then regenerated and recycled.
To provide intimate contact between the solvent and the oil, packed or trayed
towers are used, shown in Fig. 13.7. Newer extraction units are rotating disk contactor
(RDC) or centrifugal extractors for smaller volumes needed for extraction.
Today, phenol, furfural, N-methyl-2-pyrtolidone (NMP) and cresylic acid are
widely used as solvents. Liquid sulfur dioxide, chlorinated ethers, and nitrobenzene
also have been used. Table 13.3 compares the properties of N-methyl-2-pyrrolidone
(NMP), furfural and phenol.
Furfural is the most widely used solvent for lubricant oil manufacture. In furfural
extraction, the feed is introduced into a continuous countercurrent extractor at a
temperature that is a function of viscosity. The temperature at the top of tower is
below the miscibility temperature between furfural and oil, usually in the range of
220–300 °F (104–149 °C). The gradient from top to bottom of the tower is between
60–90 °F (33–50 °C), hot at the top. The furfural-dispersed phase flows downward
through the continuous oil phase. Extract is recycled at a ratio of 0.5:1. Furfural-
to-oil ratios range from 2:1 for light stocks to 4.5:1 for heavy stocks. Furfural is
easily oxidized and polymerized, so it must be protected with inert gas and carefully
controlled heat exchange. A complex solvent recovery system is needed due to the
formation of an azeotrope between furfural and water. The most important operation
variables are the furfural-to-oil ratio, extraction temperature and extract cycle ratio.
The phenol process was first commercialized in 1930 and has been extensively used
ever since, being the second only to the furfural process. Phenol is much easier to
recover than furfural. Although phenol has high solvent power and low cost, its
toxicity necessitates special handling facilities.
In phenol extraction, phenol is introduced at the top of tower at higher temperature
than the oil coming up from the bottom. The oil coming out of phenol-rich phase
at lower portion of the tower reverses direction and rises to the top as reflux. The
temperature at the top is below the miscible temperature of the mixture, while the
temperature at the bottom is maintained at 20 °F (11 °C) lower than the top. The
most important operating parameters are phenol-to-oil ratio (treat rate), extraction
temperature, and percentage of water in phenol. Treat rates range from 1:1 to 2.5:1
depending on the quality and viscosity of the feed and the quality of product desired.
NMP is a weak base and is an aprotic solvent with low volatility and toxicity. In the
late 1970s, NMP was offered as a selective solvent for lube oil refining by Texaco
and Exxon. It was developed to replace phenol for safety, health, and environmental
concerns. Compared to phenol NMP boils at 40 °F (22 °C) higher, has a 115 °F
(64 °C) lower melting point, and a 69% lower viscosity at 122 °F (50 °C). NMP is
completely miscible and forms no azeotrope with water. As shown in Fig. 13.8, a
portion of distillate or deasphalted oil feed is used as the lean oil in an absorption
tower to remove the NMP from exiting stripping steam. The rich oil is combined
266 13 Lubricant Processes and Synthetic Lubricants
with the remainder of the feed and heated to a desired temperature before going into
the treat tower. The raffinate is the desired product, while the extract is enriched with
aromatics.
NMP has higher recovery rate than phenol, with losses less than half of that of
phenol. It also allows greater throughput for a given size tower. The treat rate and
product quality are the same as for phenol extraction. NMP is better than furfural
and phenol in terms of VI improvement, whereas furfural performs better than NMP
and phenol in terms of raffinate yield.
13.11 Dewaxing
Heavy n-paraffins are waxy. They solidify in cold weather, so during lube base stock
production, they must be removed, either by extraction, conversion, or isomeriza-
tion. The dewaxing of lubrication oil represents the largest use of scraped-surface
continuous crystallizers.
The dewaxing technologies in commercial units today are solvent dewaxing and
catalytic dewaxing. Figure 13.9 shows solvent and catalytic dewaxing processes in
lubricant base stock manufacturing [3]. Note that solvent dewaxing produces slack
wax as a co-product, but slack wax is not produced in catalytic dewaxing. Solvent
dewaxing processes are based on a sequence of distillation and solvent extraction,
13.11 Dewaxing 267
Fig. 13.9 Solvent and catalytic dewaxing processes for lubricant base stock manufacturing [4]
Solvent dewaxing removes wax (n-paraffins > C20 ). The main process steps include
mixing the raffinate with the solvent, chilling the mixture to crystallize wax, and
recovering the solvent, shown in Fig. 13.10. The solvent reduces viscosity for filtra-
tion. The polarity of the solvent-oil mixture increases, decreasing the solubility of
wax in oil and promoting the formation of more-compact wax crystals. Commonly
used solvents include methyl ethyl ketone (MEK), methyl isobutyl ketone (MIBK),
MEK/MIBK, toluene, MEK/toluene or propane. MEK/MIBK refrigeration require-
ments are lower than for MEK/toluene due to lower wax solubility.
Both dewaxed oil (DWO) and wax are valuable products. For use as lube
base stock, important DWO properties include pour point, cloud point and low-
temperature fluidity. For wax, the properties of interest are oil content, melting point
and needle penetration depth.
268 13 Lubricant Processes and Synthetic Lubricants
Propane dewaxing uses liquid propane as the solvent. The vessels in the unit must
be able to operate at elevated pressure. Figure 13.11 shows a propane dewaxing
plant. The feed is mixed with a dewaxing aid and passes through a pre-chiller (heat
exchanger) before entering the chillers. Propane is used both as a refrigerant and a
diluent for direct chilling. The process requires careful pressure control. The chilled
solution is sent to a filter to separate wax from oil. Both the filtrate and wax go
through separate sets of heat exchangers to a propane recovery unit.
before and after the SSE to reduce the slurry viscosity and enhance heat transfer.
After exiting the SSC, the cold slurry enters the continuously rotating filter feed
drum, where the wax crystals are filtered under vacuum. Wax is scraped off the
filter surface and sent to the wax recovery section. The solvent/filtrate is sent to the
shell side of the SSE on its way to the dewaxed oil (DWO) recovery section. The
filtrate may also be recycled back to the slurry to adjust the final dilution ratio before
filtration.
Table 13.4 compares the advantages and disadvantages of propane and ketones for
dewaxing. Propane is readily available, less expensive and easier to recover. Propane
dewaxing uses direct chilling to reduce capital and maintenance costs. It has high
filtration rates and is accompanied by resin and asphaltene rejection. There is a
large difference between the filtration temperature and pour point of the dewaxed
oils (25–45 °F). Propane dewaxing requires the use of dewaxing aides to get good
filtration rates, and it requires higher pressure equipment.
For ketone dewaxing, there is a small difference between the filtration temperature
and pour point of dewaxed oil (9–18 °F). Hence, it has lower pour point reduction
capability. But the process has greater heat efficiency and lower refrigeration require-
270 13 Lubricant Processes and Synthetic Lubricants
ments. The chilling rate is fast, and the filtration rate is good (but lower than with
propane).
Until recently, the improvements in lubricant oils for passenger and commercial vehi-
cles were largely achieved by the use of better additives, such as antioxidants, anti-
wear agents, and viscosity improvers. However, additives alone have not been able to
meet the fuel economy and performance requirements for newer, improved vehicles.
Other than synthetic lubricants, which will be discussed later, several new technolo-
gies, including catalytic dewaxing, have been developed to meet such demands.
Most catalytic dewaxing processes involve hydroprocessing or hydroconver-
sion [4, 5]. Two main processes are (1) hydrocracking to break down large wax
272 13 Lubricant Processes and Synthetic Lubricants
Fig. 13.14 Comparison of wax crystals by Dilchill and conventional dewaxing [3]
The Mobil Lube Dewaxing (MLDW) process was developed in the mid-1970s as
an alternative to solvent dewaxing. The MLDW catalyst uses medium pore ZSM-
5, a catalytic shape-selective catalyst discovered by Mobil in the late 1960s. This
zeolite selectively cracks the wax molecules on the acid sites in the pores while
preserving the valuable base stock molecules. The ZSM-5 is shape selective. The
constrictive, intersecting two-dimensional pore geometry on the order of 5 Å allows
linear paraffins and linear portion of slightly branched isoparaffins to enter the pores to
be cracked into smaller hydrocarbons over acid sites, while rejecting highly branched
isoparaffins, naphthenes (cycloparaffins), and aromatics. Large sulfur and nitrogen
species cannot get deep inside the ZSM-5 pores to deactivate the zeolite. Therefore,
it is robust for a full range of raffinates. The cracking products include propane,
butane, naphtha, and middle distillates with lower carbon numbers.
The uncracked material has a much lower pour point and oxidative stability.
The desired reaction conditions are determined by the feed and product objectives.
Temperatures range from 560 to 700 °F (293–371 °C) under a wide range of hydrogen
partial pressures: from 300 to 2000 psig (20–180 bar). Gas rates can range from
500–5000 scf/bbl.
13.11 Dewaxing 273
Compared to solvent dewaxing, ZSM-5 removes more wax when achieving the
same pour point on a common feed, with lower yield. The cracking produces a small
amount of olefins that must be removed by hydrofinishing.
of 0.6 nm. The pore geometry allows long-chain wax molecules to enter the cata-
lyst pores while excluding isoparaffins, which have a larger effective dimeter due to
branching. The acidity of the catalytic sites is less than in ZSM-5. This results in less
cracking and greater preservation of molecules in the base oil range. At the noble
metal sites, intermediate iso-olefins are rapidly saturated. In addition to preventing
cracking, this prevents coke formation and maintains catalyst activity and stability.
Final stabilization is achieved in the hydrofinishing unit.
Several processes improve the color, odor, thermal and oxidative stability, and demul-
sibility of dewaxed oil (DWO) for use as lube base stock, including selective hydro-
cracking, hydrofining and clay contacting.
Hydrofinishing is a mild hydrotreating process using Co-Mo catalysts to remove
sulfur, nitrogen and oxygen compounds; such compounds increase the color and
decrease the color stability of lube oils. In conventional hydrofinishing, reaction
temperatures range from 400 to 650 °F (204–343 °C) under 500–800 psig (34–55 bar)
pressure. The LHSV is 0.5–2.0 v/h/v, and hydrogen gas rate is 3–4 times consumption.
In certain hydrofinishing processes, operating at higher pressures (up to 3000 psig,
200 bar) and lower temperatures provides essentially complete removal of aromatics
in addition to hetero-atom removal.
Clay Contacting contacts DWO with activated clay, which adsorbs aromatic,
sulfur and nitrogen compounds to improve stability of the oil.
13.13 GTL (Fischer-Tropsch) Process 275
(syngas)
Fig. 13.16 Gas-to-liquid lubricant base stock through Fischer-Tropsch process [4]
(2n + 1) H2 + n CO → Cn H(2n+2) + n H2 O
where n is an integer.
Recent advances in catalysts and process design have enabled the commercializa-
tion of gas-to-liquid (GTL) processes through F-T chemistry. As shown in Fig. 13.16,
natural gas or methane is converted into a mixture of carbon monoxide and hydrogen,
often called synthesis gas or syngas [4]. Then the syngas is converted into a mixture
of linear paraffins and some oxygenates. In most cases, GTL process conditions are
adjusted to maximize diesel. However, by-products with higher (and lower) carbon
numbers also are generated. C20 + GTL products require considerable hydrodeoxy-
genation and selective isomerization of linear paraffins into isoparaffins to make them
suitable as lubricant base stocks.
13.14 Wax
The raw wax from dewaxing units, called slack wax, is soft because it still contains
some oil. The oil must be removed to convert slack wax into hard wax the meets
“food grade” specifications for oil content, melting point, and needle penetration.
Methyl isobutyl ketone (MIBK) is used in a wax deoiling process to prepare food-
grade wax, Food-grade wax serves as a coating for milk cartons and other packaging.
It is a component of certain medicines, and it is blended directly into foods, such as
“construction chocolates,” to provide hardness and/or modify melting properties. In
recrystallization deoiling, the wax crystals from the dewaxing plant are melted in a
heat exchanger in the presence of a lean solvent. The wax is then recrystallized in
SSEs and SSCs to get rid of excess oil. Wax can be fractionated by consecutive slow
cooling and filtration at various temperatures, with collection of wax crystals at each
specific temperature.
276 13 Lubricant Processes and Synthetic Lubricants
Petroleum wax is of two general types: paraffin wax from petroleum distillates and
microcrystalline wax from residua. Table 13.5 compares waxes from these different
sources.
There are many forms of finished lubricant oil products and specialty products for
a wide variety of applications. Finished Lube Oils include engine oils, transmission
fluids, gear oils, turbine oils, hydraulic oils, metal working (cutting) oils, greases,
paper machine oils, etc. Specialty products include white oils for foods, pharmaceu-
ticals and cosmetics; agricultural oils such as those used as orchard sprays; electrical
oils for electrical transformers (heat transfer media), etc.
In the early 1950s, extreme demands for high- and low-temperature lubricity in jet
engines could not met by mineral lube oils. Polyol esters provide superior thermal
oxidative stability, lubricity and volatility. During oil drilling in Alaska in the mid-
1960s, mineral lube oils solidified under severe Alaskan cold weather conditions and
could not function. It was found that poly alpha olefins (PAO), initially studied by
Socony-Mobil in the early 1950s, provide excellent low temperature flow properties.
In 1973, Mobil introduced PAO-based SHC in Europe, followed by fuel saving SAE
5W-20 Mobil 1 in the U.S.
Synthetic lube base stocks are more expensive, but they are preferred: (1) when
conventional lubes cannot meet specific performance demands, (2) when a synthetic
lube can offer economic benefits for overall operations, and (3) for modern machines
and equipment that are operated under increasingly severe conditions, where it is
13.16 Synthetic Lubricant Oils (Groups IV and V Oils) 277
Fig. 13.17 Synthetic lube base stocks derived from petrochemicals [7]
278 13 Lubricant Processes and Synthetic Lubricants
The remaining PAOs have viscosities of 10–100 cSt. Among the PAO of different
carbon numbers synthesized from ethylene, 1-decene is the most commonly used
starting material. It can produce higher oligomers, shown in Fig. 13.18, depending
on the catalyst, BF3 or AlCl3 , and reaction conditions in a multistage, continuously
stirred tank reactor (CSTR). Table 13.6 lists some properties—kinematic viscosity,
viscosity index and pour point—of various PAO oligomers synthesized from different
linear 1-olefins. The combination of C30 and C40 provides a very narrow molecular
weight distribution for a 4 cSt oil, with less lower carbon number components. This
is important because the smaller components can lower flash point by degrading oil
volatility.
PAO with a given carbon number is a mixture of many isomers with different types
of branching. This is because the use of BF3 or AlCl3 to catalyze oligomerization of
LAO via carbonium ion intermediates can have the charge center moved along the
hydrocarbon chain, resulting in non-uniform connections between the monomers,
shown in Figure 13-19. The irregular branching is beneficial for very low poor point
properties of PAOs. Using metallocene as a catalyst, the oligomerization of LAO
produces oligomers with a uniform comb-like structure shown in Fig. 13.20. Thus, it
can produce high molecular weight PAOs with narrow molecular weight distributions,
with higher VI, wider viscosity range, and lower pour points.
PAOs do not contain ring hydrocarbons, naphthenes and aromatics, which gives
them intrinsic oxidative stability compared to conventional Groups I, II and III min-
Table 13.6 Properties of PAO derived from different linear 1-olefins [7]
Name Carbon number Kinematic viscosity cSt at Viscosity index Pour
100 °C 40 °C −40 °C point
(°C)
Propylene C30 7.3 62.3 >99,000 70 –
decamers
Hexane C30 3.8 18.1 7850 96 –
pentamers
Octene tetramers C32 4.1 20.0 4750 196 –
Decene trimers C30 3.7 15.6 2070 122 <−55
Undecene C33 4.4 20.2 3350 131 <−55
trimers
Dodecene C36 5.1 24.3 13,300 144 −45
trimers
Decene C40 5.7 29.0 7475 141 <−55
tetramers
Octene C40 5.6 30.9 10,225 124 –
pentamers
Tetradecene C42 6.7 33.8 Solid 157 −20
trimers
H+ H2 /catalyst
from
BF3
catalyst
Fig. 13.19 Formation of PAO via carbonium ion intermediates using BF3 or AlCl3 as a catalyst
eral oil base stocks. PAOs also have superior viscometric properties compared to the
mineral oil base stocks, with lower volatility due to narrow molecular weight ranges.
The largest volume of PAO is made for premium automotive engine oils, which
provide improved engine protection, extended oil drain interval with reduce oil con-
sumption, improved fuel economy, excellent low-temperature fluidity and pumpa-
bility, and high temperature oxidative resistance. Figure 13.19 compares viscosity
changes in engine tests of PAO with Group I/II and Group III oils [7]. After over
200-h operations, the mineral oils become too thick for viscosity measurement, while
PAO is still in a suitable operation range. The high viscosity PAOs are used as indus-
trial oils and greases, with longer fatigue life, a wider operating range due to higher
280 13 Lubricant Processes and Synthetic Lubricants
Ester based lubricants of natural sources, such as lard and vegetable oil, have been
known throughout human history. However, demands for synthetic ester lubricants
started after World War II to meet jet engine requirements. Esters are made from
the reaction of acids with alcohols. Among many types of ester lubricants, dibasic
esters, polyol esters and aromatic esters are the most commonly used. A comparison
of their properties is shown in Table 13.7.
Fig. 13.20 Comparison of PAO structures by metallocene catalyst versus conventional BF3 or
AlCl3 as catalyst
Ester base stocks have lower volatility than PAO and mineral oil base stocks of
comparable viscosities because of their higher polarity. The general rank of volatility
is PE ester < TMP ester < dibasic ester < PAO Group I or II mineral oils. Dibasic
and polyol esters have excellent biodegradability. Aromatic esters in general are less
biodegradable than aliphatic dibasic and polyol esters.
Many commonly used additives are more soluble in ester base stocks than in min-
eral oil base stocks. The high solubility of organic acids and sludge during services
makes low viscosity ester fluids ideal for co-base stocks of PAO for solvency and
dispersancy. A major issue for ester fluids is the formation of corrosive acids when
contaminated with water. Esters made from aromatic and other sterically hindered
acids improve hydrolytic stability. Ester base stocks provide mild film protection at
lower temperatures. Higher molecular weight acids, upon decomposition, are bound
to metal to provide some degree of wear protection and friction reduction.
Dibasic esters
Adipate Iso- 5.4 27 139 −51 4.8 92
C13 H27
Sebacate Iso- 6.7 36.7 141 −52 3.7 80
C13 H27
Polyol esters
n-C8 /n-C10 PE 5.9 30 145 −4 0.9 100
n-C5 /n- PE 5.9 33.7 110 −46 2.2 69
C7 /iso-C9
n-C8 /n-C10 TMP 4.5 20.4 137 −43 2.9 96
Iso-C9 TMP 7.2 51.7 98 −32 6.7 7
n-C9 NPG 2.6 8.6 145 −55 31.2 97
Aromatic esters
Phthalate Iso- 8.2 80.5 56 −43 2.6 46
C13 H27
Phthalate Iso-C9 5.3 38.5 50 −44 11.7 53
Trimellitate Iso- 20.4 305 76 −9 1.6 9
C13 H27
Trimellitate n-C7 /n- 7.3 46.8 108 −45 0.9 69
C9
282 13 Lubricant Processes and Synthetic Lubricants
Dibasic esters are made from the reaction of carboxylic diacids with alcohols. The
most commonly used diacid is adipic (hexanedioic acid), usually combined with
ethylhexanol or isotridecanols to give balanced high VI and low temperature prop-
erties (see Table 13.7). Dibasic esters are often used with PAO as a co-base stock to
improve solvency with additives and swell properties of final formulated lubricant
products.
Polyol esters are made from reaction of monoacids with polyols. The most commonly
used polyols are pentaerythriol (PE), trimethylolpropane (TMP) and neopenttylgly-
col (NPG), shown in Fig. 13.22. High VI and low pour point base stocks can be made
by careful choice of acids with branching.
A polyol ester molecule lacks β-H adjacent to carbonyl oxygen, so it cannot
undergo low energy β-H transfer to decompose into olefin and acid as dibasic esters,
shown in Fig. 13.23. Hence, it can only be decomposed by C-O or C-C cleavage
into radicals at extremely high temperatures. Therefore, polyol esters are thermally
stable up to 250 ºC.7 Polyol esters exhibit thermal stability in the following order:
PE esters > TMP esters > NPG esters.
The reaction between phthalic anhydride or trimellitic anhydride with alcohols results
in aromatic esters. They have relatively low VI’s, as shown in Tables 13.4, 13.5 and
13.6, and only are used in special industrial oil applications.
13.16 Synthetic Lubricant Oils (Groups IV and V Oils) 283
EO-based PAGs are typically waxy and have poor low-temperature properties.
They are typically used to formulate water-based lubricants, especially fire-resistant
hydraulic oils. PO-based PAGs are excellent lubricant oils with high VI and low
pour point. They are less soluble to water than EO-based PAGs, but are not oil mis-
cible. EO/PO-based PAGs have a better combination of VI and low pour points than
PO-based PAGs. The are used as base stock in industrial circulation/bearing/gear
oils.
References
1. 2016 Annual Book of ASTM Standards (2016) Section 5: petroleum products, liquid fuels,
and lubricants, vols 05.01 and 05.02. American Society for Testing and Materials (ASTM)
International, West Conshohocken, PA
2. Hsu CS (ed) (2003) Analytical advances for hydrocarbon research. Kluwer Academic/Plenum
Publishers, New York
3. (a) Beasley BE (2006) Conventional lube base stock manufacturing. In: Hsu CS, Robinson PR
(eds) Practical advances in petroleum processing. Springer, New York (Chapter 15). (b) Beasley
BE (2017) Conventional lube base stock. In: Hsu CS, Robinson PR (eds) Springer handbook of
petroleum technology. Springer, New York (Chapter 33)
4. Cody IA (2006) Selective hydroprocessing for new lubricant standards. In: Hsu CS, Robinson
PR (eds) Practical advances in petroleum processing. Springer, New York (Chapter 16)
References 285
5. Lee SK, Rosenbaum JM, Hao Y, Lei GD (2017) Premium lubricant base stocks by hydropro-
cessing. In: Hsu CS, Robinson PR (eds) Springer handbook of petroleum technology. Springer,
New York (Chapter 34)
6. Lei G-D, Dahlberg A, Krishna K (2015) All hydroprocessing route to high quality lubricant
base oil manufacturing using Chevron ISODEWAXING Technology. http://www.nt.ntnu.no/
users/skoge/prost/proceedings/aiche-2008/data/papers/P138184.pdf. Accessed 15 Aug 2015
7. (a) Wu MM, Ho SC, Forbus TR (2006) Synthetic lubricant base stock processes and products.
In: Hsu CS, Robinson PR (eds) Practical advances in petroleum processing. Springer, New York
(Chapter 17). (b) Wu MM, Ho SC, Luo S (2017) Synthetic lubricant base stock. In: Hsu CS,
Robinson PR (eds) Springer handbook of petroleum technology. New York: Springer, 2017
(Chapter 35)
Chapter 14
Other Refining Processes
(816 °C). The product of the initial reaction is a mixture of H2 , CO, CO2 , residual
methane, and in some cases traces of other hydrocarbons. The initial product goes
to one or more water shift reactors, where the shift reaction between H2 O and CO
yields CO2 and additional hydrogen.
CH4 → 2H2 + C
C + H2 O → CO + H2
2CH4 + O2 → CO + 4H2
Methanation: CO + H2 → CH4
Fig. 14.1 Hydrogen plant with a methanizer [1] HTSC: high-temperature shift converter; LTSC:
low-temperature shift converter
Natural gas can contain significant amounts of hydrogen sulfide. Some gas processing
plants in Alberta, Canada, recover more than 2000 tons of sulfur per day using
the Claus process. Sulfur can also be recovered by processes such as LOCAT and
Selectox. A small Selectox unit converts H2 S to sulfur on Platform Irene off the coast
of California [2].
290 14 Other Refining Processes
Fig. 14.2 Hydrogen purification by pressure-swing adsorption (PSA) [1] HTSC: high-temperature
shift converter
Flue-gas scrubbing is a refiner’s last chance to keep NOx and SOx out of the air.
In wet flue-gas desulfurization, gas streams containing SOx react with an aqueous
slurry containing calcium hydroxide Ca(OH)2 and calcium carbonate CaCO3 . Reac-
tion products include calcium sulfite (CaSO3 ) and calcium sulfate (CaSO4 ), which
precipitate from the solution. NOx removal is more difficult. Wet flue-gas scrub-
bing removes about 20% of the NOx from a typical FCC flue gas. To remove the
rest, chemical reducing agents are used. In the Selective Catalytic Reduction (SCR)
process, anhydrous ammonia is injected into the flue gas as it passes through a bed
of catalyst at 500–950 °F (260–510 °C). The chemical reaction between NOx and
ammonia produces N2 and H2 O.
When sulfur-containing feeds pass through hydrotreaters or conversion units, most
of the sulfur is converted into H2 S, which eventually ends up in off-gas streams.
Amine absorbers remove the H2 S, leaving only 10–20 wppm in the treated gas
streams. H2 S is steam-stripped from the amines, which are returned to the absorbers.
The H2 S goes to the refinery sulfur plant.
Amine gas treating, also known as amine scrubbing, gas sweetening and acid
gas removal, refers to a group of processes that use aqueous solutions of vari-
ous alkanolamines such as monoethanolamine (MEA), diethanolamine (DEA), tri-
ethanolamine (TEA), methyldiethanolamine (MDEA), dipropanolamine (DIPA) and
diglycolamine (DGA) to remove hydrogen sulfide and carbon dioxide from gases
and hydrocarbon streams. Amine scrubbing is common in refineries, and is also used
in petrochemical plants, natural gas processing plants and other industries.
The treating chemistry at low partial pressure of the acid gas is:
2RNH2 + H2 S → (RNH3 )2 S
2RNH2 + CO2 + H2 O → (RNH3 )2 CO3
(RNH3 )2 S + H2 S → 2RNH3 HS
(RNH3 )2 CO3 + H2 O → 2RNH3 HCO3
292 14 Other Refining Processes
As shown in Fig. 14.3, sour gas containing carbon dioxide and/or hydrogen sulfide
is charged to a gas absorption tower or liquid contactor where the acid contaminants
are absorbed by counter flowing amine solutions (i.e. MEA, DEA, MDEA). The
stripped sweet gas is removed overhead, and the amine is sent to a regenerator. In
the regenerator, the acidic components are stripped by heat and reboiling action and
routed to recovery units. The amine is recycled.
Claus sulfur-recovery units, shown in Fig. 14.4, burn hydrogen sulfide in enough
air to form a mixture of H2 S and SO2 in a 2:1 molar ratio. In downstream beds of
alumina catalyst, H2 S reacts with SO2 to form elemental sulfur and water.
In a Claus unit with three catalyst beds, as shown in the figure, overall H2 S recovery
in less than 98%. Due to the high concentration of H2 S, the tail gas is toxic and must
be incinerated to SO2 or routed to a tail-gas treating unit (TGTU). Incineration occurs
in a high stack to disperse the SO2 . After incineration, no H2 S remains.
14.2 Sulfur Removal and Recovery 293
In most of the world, air pollution regulations limit SOx in stack emissions to 10
ppm. In some locales, the limit is as high as 100 ppm. Therefore, tail gas treatment
is a necessity for Claus units.
At the sulfur plant, H2 S is combined with sour-water stripper off-gas and sent to
a Claus unit. Almost every refinery in the world uses some version of this process
to convert H2 S into elemental sulfur. H2 S and a carefully controlled amount of air
are mixed and sent to a burner, where about 33% of the H2 S is converted to SO2 and
water. In several units, the combustion air is enriched with oxygen to increase plant
capacity.
From the burner, the hot gases go to a reaction chamber, where the reactants and
products reach equilibrium. Elemental sulfur is produced by the reversible reaction
between SO2 and H2 S. Ammonia comes in with the sour-water stripper off-gas. In
the Claus process, it is thermally decomposed into nitrogen and water.
In the Claus burner, combustion temperatures reach 2200 °F (1200 °C). Much of
the heat is recovered in a waste-heat boiler, which generates steam as it drops the
temperature to 700 °F (370 °C). Next, the process gas goes to a condenser, where
it is cooled to about 450 °F (232 °C). At this temperature, sulfur vapors condense,
and the resulting molten sulfur flows through a drain to a heated sulfur-collection
pit. At the bottom of the drain, a seal leg maintains system pressure and keeps
unconverted gases out of the pit. Uncondensed sulfur and other gases flow to a series
of catalyst beds, which recover additional sulfur by promoting the reaction between
left-over H2 S and SO2 . With fresh catalyst and a stoichiometric gas composition,
the cumulative recovery of sulfur across four condensers is about 50, 80, 95, and
96–98%, respectively.
Claus tail-gas treating (TGT) units bring the total sulfur recovery up to >99.9%
in the refineries and over 99.2% in the LNG plants depending feed gas composition
[4]. Most tail-gas treating processes send the tail gas to the Beavon Sulfur Removal
(BSR) hydrotreater, which converts all sulfur-containing compounds (mercaptans,
294 14 Other Refining Processes
SO2 , SO3 , COS, CS2 and various vapor forms of Sx ) into H2 S. In the Shell Claus Off-
gas Treatment (SCOT) process offered by Shell Global Solutions, shown in Fig. 14.5,
the H2 S from the BSR section is absorbed by an amine and returned to the front of
the Claus furnace [5].
The BSR Stretford process consists of two stages [6]. In the Stretford section, the
H2 S is oxidized using an alkaline solution containing vanadium as an oxygen carrier
[7]. BSR Stretford removes essentially all of the sulfur compounds from Claus plant
tail gases to give elemental sulfur. Chemicals in the Stretford process were toxic, and
during process upsets, the sulfur was contaminated by vanadium and could not be
sold without purification. Hence, the process is no longer used.
For all sulfur-recovery processes, vapor containing the last traces of unrecovered
sulfur go to an incinerator, where they are converted into SO2 and dispersed into the
atmosphere.
At less than about 20% H2 S in acid gas or tail gas mixtures, a stable flame cannot
be maintained in a Claus furnace even with a split-flow furnace design. The UOP
Selectox process catalytically converts H2 S to elemental sulfur without the need for
reaction furnace. It is advantageous for recovering high-purity sulfur from acid gas
feed streams containing from less than 5% to about 20% H2 S [9]. As mentioned
above, the process has even been used for natural gas processing on an offshore
platform. [2] The feed gases are passed over a fixed bed of a proprietary Selectox
14.2 Sulfur Removal and Recovery 295
catalyst at 160–370 °C. It achieves up to 97% sulfur recovery at the start of a catalyst
cycle, and is often used for tail gas treatment in conjunction with the BSR process
discussed above for overall recovery >99.9%.
In the LO-CAT® process, offered by Merichem, H2 S is air-oxidized to sulfur in
an aqueous solution containing a dual chelated iron catalyst under mildly caustic
conditions. The system is designed with separate absorber and oxidizer vessels. The
absorber removes the H2 S from sour gas, converting it to elemental sulfur. The
oxidizer regenerates the catalyst which is pumped between the vessels. It achieves
sulfur removal efficiencies up to >99.9% in many different applications and industiies
[10].
LO-CAT can be applied to all types of gas streams, including air, natural gas, CO2 ,
amine acid gas, biogas, landfill gas, refinery fuel gas, etc. The liquid catalyst adapts
to the variation in flow and concentration easily. Units require minimal operator
attention and are well-suited to removing H2 S from sour gas streams.
Selective catalytic reduction (SCR) [11] removes nitrogen oxides (NOx ) by reaction
with ammonia to produce nitrogen via the following main reactions.
2SO2 + O2 → 2SO3
2NH3 + SO3 + H2 O → (NH4 )2 SO4
NH3 + SO3 + H2 O → NH4 HSO4
In the process, the gas stream containing NOx , such as flue gas or exhaust gas, is
mixed with a gaseous reductant, typically anhydrous ammonia, aqueous ammonia,
or urea, and is adsorbed onto a catalyst. SCR catalysts are made from various ceramic
carriers, such as titanium oxide, and active catalytic components of oxides of base
metals (such as vanadium, molybdenum, and tungsten), zeolites, or various precious
metals. Nitrogen oxides are reduced to nitrogen. If urea is used as the reductant, CO2
is also produced.
296 14 Other Refining Processes
Conventional crude oil is getting harder to find, despite the fact that estimates of
proven reserves increased substantially with the recent advent of “fracking” technol-
ogy to produce oil and gas from tight formations and non-conventional reservoirs.
Fracking oils, oil sand bitumen, and extraheavy crude oils are called unconventional
crude oils.
The primary use of oil and gas is energy production. Declining availability of
conventional petroleum can be offset to some extent by non-fossil sources. But
unconventional fossil fuels are likely to carry most of the load. Liquid fuels will
be produced from natural gas, bitumen, kerogen and coal. Some of these contain
daunting amounts of contaminants. Compared to conventional oil, converting these
materials will be costly and more difficult. But the technology for doing so already
exists.
14.5 Alternatives to Petroleum 297
14.5.1 Biomass
Biomass is organic matter derived from living or recently living organisms, such as
plants, algae, grass, vegetables, animal fats, etc. It can be combusted to produce heat
directly or converted into biofuels, most commonly ethanol as a blend component
or alternative to gasoline, and as triglycerides as biodiesels. Prior to use in diesel
engines, cellulose must be converted into a suitable liquid, and vegetable oil must be
stabilized, usually by hydrogenation to remove oxygenates.
In Brazil, ethanol produced from sugar cane replaces a huge percentage of that
country’s liquid transportation fuel. Growing sugar cane impacts land use, but the
lifecycle energy gain is about 8:1.
The United States now requires the blending of ethanol into gasoline. The require-
ment was established to decrease air pollution and reduce dependence on imported
oil. In practice, almost all US ethanol is produced from corn, which is a poor sub-
strate when compared to sugar cane. The estimated energy gain for ethanol from
corn ranges is from 1.1 to 1.5. US growers depend heavily on irrigation, which is
depleting the Ogallala aquifers and other important sources of ground water. U.S.
corn growers consume tremendous amounts of fertilizer, which increases emissions
of greenhouse gases, particularly CO2 and N2 O. Using corn for fuel has increased
298 14 Other Refining Processes
the price of food nationwide [13–15] Adding 10% ethanol to gasoline decreases the
demand for petroleum by less than 10%. However, the gas mileage is less because
the heat content of ethanol is lower than the heat content of hydrocarbons. Ethanol
increases the RVP of gasoline, thereby increasing emission of light hydrocarbons.
Many countries offer incentives to add oils from plants to diesel fuels. Bio-oils
that are derived from pyrolysis of biomass are not suitable for direct use in existing
engines due to their high levels of oxygen, immiscibility with fossil fuels and high
tendency to polymerize when exposed to air. So they are mixed in small amounts with
conventional petroleum to be hydroprocessed in existing refineries. Given enough
time, we could develop a new biomass-based chemical industry [16].
The biomass used to generate biofuels can cover a broad range of materials that
include: food crops (1st generation biofuels), non-food crops (energy crops) or non-
edible portions of food crops (2nd generation biofuels) and microalgae (3rd genera-
tion biofuels) [17, 18]. The 4th generation of biofuels aims to combine production of
biofuels with capturing and storing CO2 (CCS) so that it would be carbon negative
rather than carbon neutral with respect to air.
The 1st generation biofuels have achieved a certain degree of success in com-
mercialization, especially under government subsidies or mandates. The bioethanol
derived from corn and sugarcane, for example, is unable to replace gasoline com-
pletely without major modification to the internal combustion engine, and con-
sequently, it must be blended with gasoline in order to be used within existing
engines. Biobutanol, however, can be a good candidate for a gasoline alternative
in existing engines.
Many of the energy production and utilization cycles based on lignocellulosic
biomass have low or near-zero greenhouse gas (GHG) emissions on a life-cycle
basis. A much overlooked concept that should increase the popularity of utilizing
these biomass feedstocks is the idea of use within a fully integrated biorefinery, much
like what has already been developed within petroleum refining.
One kind of biocrude is fast pyrolysis oil from wood or other biomasses, including
municipal wastes that otherwise might go to a sanitary landfill. It has high oxygen
content, mostly in the forms of ketones and esters. Intermediates such as alcohols,
carboxylic acids or aldehydes may be formed from carbonyl groups upon hydro-
genation. Consequently, the carbonyl groups are expected to follow one of the three
deoxygenation reactions: hydrodeoxygenation (HDO) of alcohols, decarboxylation
of acids and decarbonylaton of aldehydes, to produce pure hydrocarbons. Pyrolysis
oil (bio-oil) could serve as a source of certain chemicals.
Biocrude and other biomass oils, such as vegetable oils, can be blended with
high acid crude, heavy crudes, or distillation bottoms, to be co-processed. The ideal
bio-oil upgrading catalysts need high activity for deoxygenation and must be able to
withstand large quantities of coke. The issue of coke can be handled by employing
14.5 Alternatives to Petroleum 299
References
1. Crew MA, Shumake BG (2017) Hydrogen production. In: Hsu CS, Robinson PR (eds) Springer
handbook of petroleum technology. Springer, New York (2017) (Chapter 24)
2. Bertram RV, Robinson PR (1989) Selectox process for sulfur recovery offshore. In: The 68th
annual gas processors association meeting. San Antonio, TX, 14 Mar 1989
3. Robinson PR (2017) Sulfur removal and recovery. In: Hsu CS, Robinson PR (eds) Springer
handbook of petroleum technology. Springer, New York (2017) (Chapter 20)
4. Tail Gas Treating: TGT. https://www.chiyoda-corp.com/technology/en/upstream_
gasprocessing/tail_gas_treating_tgt.html. Accessed 6 Aug 2014
5. SCOT. http://www.shell.com/business-customers/global-solutions/gas-processing-licensing/
gas-processing-technolgies-portfolio/_jcr_content/par/textimage.stream/1444035481012/
824ffff53bbcedc4da5daa9094ec8cf1a51451ce5fd2ca8f3021a44b6d2735fb/factsheet-scot-
screen.pdf. Accessed 6 Aug 2014
6. Fenton DM, Gowdy HW (1979) The chemistry of the Beavon sulfur removal process. Environ
Int 2(3):183–186
7. Stretford process. https://en.wikipedia.org/wiki/Stretford_process
8. LaRue K, Grigson SG, Hudson H (2013) Sulfur plant configuration for weird acid gases. In:
Presented at 2013 Laurence Reid gas conditioning conference. The University of Oklahoma,
24–27 Feb 2013. http://www.ortloff.com/files/papers/LRGCC2013.pdf. Accessed 24 Nov 2016
9. http://www.ogj.com/articles/print/volume-99/issue-35/processing/long-term-operating-data-
shed-light-on-selectox-process.html Accessed 24 Nov 2016
10. Merichem: LO-CAT® process for cost effective desulfurization of all types of gas streams.
http://www.merichem.com/images/casestudies/Desulfurization.pdf. Accessed 6 Aug 2014
300 14 Other Refining Processes
Natural gas comes from gas wells, oil wells, and condensate wells. Natural gas at the
wellhead contains primarily methane (70–90%) with impurities. Natural gas from
oil wells is called associated gas. It can be free gas or dissolved in the oil. Natural
gas from a condensate well is produced along with volatile hydrocarbon condensate.
In addition to methane, natural gas mixtures contain ethane, propane, butane
and pentanes, water vapor, hydrogen sulfide, mercaptans, carbonyl sulfide, carbon
dioxide, ammonia, helium, nitrogen and other components. Natural gas is the major
commercial source of helium which is present in natural gas as an inert gas with
nitrogen. Helium is thought to form from radioactive decay (α-decay) of uranium
and thorium in granitoid rocks of Earth’s continental crust. High purity helium is
produced by the combination of cryogenic and pressure swing adsorption process.
Natural gas liquids (NGL) include ethane (35–55%), propane (20–30%), normal
butane (10–15%), isobutane (4–8%) and C5 + natural gasoline (10–15%). They are
used in enhanced oil recovery or as raw materials for oil refineries or petrochemical
plants, and sources of energy. In the U.S., hydraulic fracturing (fracking) increases
the production of natural gas as shale gas.
Natural gas produced at the wellhead, which in most cases contains contaminants
and natural gas liquids, must be processed to meet quality specifications before it can
be safely delivered to the high-pressure, long-distance pipelines and/or gas tankers
that transport the product to the consumers. Natural gas that is not within certain spe-
cific gravities, pressures, Btu (heat) content range, dew point, or water content levels
will cause operational problems, pipeline deterioration, or even pipeline rupture.
Such gas can be especially harmful to equipment in pumping stations. The water
content, as defined by water dew point, must be limited to prevent the formation
of ice and hydrate in the pipeline. The amounts of entrained hydrocarbons heavier
than ethane, as defined by hydrocarbon dew point, must be limited to prevent the
accumulation of condensable hydrocarbon liquids to block the pipeline and pumps.
Figure 15.1 shows the normal procedure to treat natural gas at the well site for
subsequent transportation to consumers. The first step is phase separation by gravity
to separate water, hydrocarbon condensates, and particulate solids from the gas in a
settler. Natural gas pretreatment typically consists of mercury removal, gas sweeten-
ing and drying. Mercury is removed by using adsorption processes based on activated
carbon or regenerable molecular sieves. Natural gas is dried by absorption in regen-
erable triethylene glycol (TEG) (glycol dehydration) or molecular sieve adsorbers
[1]. It may be necessary to remove H2 S and CO2 (acid gases) from the natural gas.
Sweet gas that contains low concentrations of sulfur compounds can be removed
by adsorption along with the removal of water, or directly sent to a gas processing
plant. For sour gas that contain high concentrations of sulfur compounds and carbon
dioxide, gas sweetening using a solvent, such as amines described in the previous
chapter, sulfinol [2] or carbonate washing shown below, is traditionally used. The
sweetened gas can then be sent to the gas processing plant or reinjected into the field
for enhanced recovery. The sulfur-rich amines solution after sulfur scrubbing is sent
to solvent recovering unit to separate the solvent from sulfur, which is sent to a sulfur
recovery plant.
If the natural gas is to be liquefied by cryogenic cooling and stored in the liquid
form, carbon dioxide must be removed by amine treatment to prevent the formation
of dry ice that would interfere with the refrigeration system and clog the pipeline.
Carbon dioxide can be sold as industrial dry ice, and in some locations it is reinjected
15.1 Natural Gas 303
into the formation. Reinjection brings the added value of reducing releases of CO2
to the atmosphere; CO2 reinjection also is known as carbon sequestration.
Natural gas does not contain olefins. Figure 15.2 shows a natural gas processing
plant to separate individual light gases [3], which is saturated gas processing plant
different than the cracked gas plant for refinery gases. The gas is compressed to
approximately 200 psig and fed to an absorber-deethanizer after phase separation,
where some ethane and lighter gases (C2 −) evolve. The oil used in the absorber-
deethanizer is usually a dehexanized naphtha with an end point of 350–380 °F. The
absorber column usually contains 20–24 trays in the top absorption section and
16–20 trays in the stripping section. The lean absorption oil is fed to the top tray
to absorb 85–90% of the C3 ’s and almost all of the C4 ’s. Significant amounts of
C2 − hydrocarbons are vaporized from the lean oil and leave the top of column with
the residue gas, from which they are recovered in a sponge absorber. The sponge
absorber usually contains 8–12 trays using kerosene or No. 2 fuel oil as the sponge
oil, which is derived as a side cut from coker fractionation or catalytic cracking
fractionation. The deethanized oil flows to the debutanizer to separate C3 + from the
absorption oil, then to the depropanizer to recover C3 from C4 ’s. The C4 stream is
sent to a deisobutanizer to separate isobutane from n-butane. During liquefaction
of natural gas, accomplished by reducing its temperature to −260 °F (−160 °C)
at atmospheric pressure, ethane can be cryogenically separated from methane or
extracted from liquefied natural gas (LNG).
304 15 Natural Gas and Petroleum Products
Table 15.2 U.S. consumption of natural gas and petroleum products (in 1000 barrels per day)
2010 2014
Natural gas liquid 2265 2448
Liquefied petroleum gas 2051 2396
Still gas (principally methane + ethane) 672 693
Ethane/ethylene 880 1048
Propane/propylene 1160 1167
Motor gasoline 8993 8921
Kero-type jet fuel 1462 1470
Distillate fuel (diesel <15 ppm S) 3211 3800
Residual fuel oil 535 257
Lubricants 131 126
Petroleum coke (marketable) 151 112
Petroleum coke (catalyst) 225 233
Asphalt 362 327
stocks. The consumptions of aviation gasoline, kerosene and wax are less than 20,000
barrels per day, and not listed in Table 15.2.
Figure 15.3 represents the variation of product distribution from a light and a heavy
crude. In this figure, gasoline and distillates represent the most valuable products.
The heavy oils would need to be converted into these higher-value products, and
would increase the costs of refining.
For petroleum products, specifications developed by American Society for Testing
Materials (ASTM) International and International Standards Organization (ISO) are
306 15 Natural Gas and Petroleum Products
widely used throughout the world. In addition to setting specifications, these institu-
tions develop and publish test methods for analyzing a wide array of materials. They
cooperate both with each other and with government regulators. For example, recent
low-sulfur gasoline and diesel directives from the U.S. Environmental Protection
Agency are incorporated by ASTM into D975 and D4814, respectively. ASTM fuel
specifications are listed in Table 15.3.
Other widely used tests and specifications are defined by licensors. For example,
UOP’s Laboratory Test Methods defines several hundred procedures for analyzing
catalysts, chemicals and fuels.
Additives are essential components of finished fuels. They increase stability,
improve flow properties and enhance performance. Cetane-improvers are routinely
added to diesel fuel, and additives that prevent intake-valve deposits are now required
in all grades of gasoline in the United States.
Petroleum gases are gases at ambient temperature and pressure. Gases produced
by refineries include methane, ethane, propane, ethylene, propylene, butylenes, and
hydrogen.
H2 S, ammonia, and CO2 also are produced in refineries as by-products. H2 S is
collected by adsorption in amine units and converted into elemental sulfur in Claus
units equipped with tail-gas recovery. Ammonia, generated from organic nitrogen,
ends up in sour water strippers and usually is destroyed in Claus units. In a few
refineries, ammonia is collected and sold. CO2 is a by-product of certain hydrogen
generation units, from which it can be collected, purified, and sold as industrial dry
ice.
Methane is the main hydrocarbon ingredient of both natural gas and refinery fuel
gas. It can be used as a refinery fuel and feedstock for hydrogen production units.
The most common of these are steam-methane reformers (SMR units), which make
high-purity hydrogen or synthetic gas (syngas), which is a well-defined mixture of
hydrogen and CO.
15.3 Petroleum Gases 307
Liquefied refinery gases are produced in the refineries from processing crude oils
and unfinished oils. They are retained in the liquid state through compression and/or
refrigeration. The reported categories are ethane/ethylene, propane/propylene, nor-
mal butane/butylene, and isobutane/isobutylene.
308 15 Natural Gas and Petroleum Products
Fig. 15.4 Motor gasoline pool from various refining processes [4]
15.4 Gasoline
Since the birth of modern automobile in 1886, the demand for gasoline to fuel
automobiles with internal combustion engines has increased dramatically. The rate
of the increase jumped in 1908, when the advent of the Model T made autos affordable
by average people. It’s estimated the number of automobiles in the world exceeded
1 billion around 2010.
The development of modern refinery processes was driven by rapid growth in
demand for transportation fuels. A variety of refining processes have been devel-
oped to produce gasoline components, including straight run distillation, isomeriza-
tion, reforming, alkylation, polymerization, catalytic cracking and hydrocracking, as
shown in Fig. 15.4. These components are in a gasoline pool for blending for meet-
ing octane rating requirements and with additives for formulated gasolines. Ninety
percent of gasoline produced in the U.S. is used as fuel in automobiles.
1. Intake stroke: the piston begins at top dead center (T.D.C.) and ends at bottom
dead center (B.D.C.). The intake valve is in the open position while the piston
pulls an air-fuel mixture into the cylinder by producing vacuum pressure into the
cylinder through its downward motion.
2. Compression stroke: the piston begins at B.D.C, or just at the end of the suction
stroke, and ends at T.D.C. The piston compresses the air-fuel mixture in prepa-
ration for ignition during the power stroke. Both the intake and exhaust valves
are closed during this stage.
3. Power stroke: The piston is at T.D.C. (the end of the compression stroke) when
the compressed air-fuel mixture is ignited by a spark plug, forcefully returning
the piston to B.D.C. This stroke produces mechanical work from the engine to
turn the crankshaft.
4. Exhaust stroke: the piston once again returns to T.D.C. from B.D.C. while the
exhaust valve is open. This action expels the spent air-fuel mixture through the
exhaust valve.
In the combustion chamber, vapor heats up during compression. Ideally, the mixture
of gasoline vapor and air in a spark-ignition engine is ignited with a spark when the
cylinder reaches a predetermined position in the cylinder. However, some compounds
can start to ignite before the piston reaches the top end (T.D.C.) near the spark
plug. This premature ignition causes engine knock, which pushes piston in wrong
direction, reduces the power of the engine, increases engine wear, and can cause
serious damage.
Gasoline is a blend of many different components. Different gasoline components
have different knock behaviors. The compression ratio (V1/V2), shown in Fig. 15.7,
is a factor in knock behavior of a gasoline mixture. Octane rating, or octane number, is
an arbitrary unit related to the smallest compression ratio at which an engine starts to
knock with a given fuel. It is based on a scale in which the octane number of n-heptane
is designated as zero and the octane number of isooctane (2,2,4-trimethylpentane) is
100. When a fuel is tested in a standard single-cylinder engine, mixtures of isooctane
and n-heptane of various percentages are used as reference standards for correlating
the knocking of the test fuel with the percentage of isooctane in the mixture.
Thus, a gasoline with the same knocking characteristics as a mixture of 94%
2,2,4-trimethylpentane and 6% n-heptane has an octane rating of 94. A rating of
94 does not mean that the gasoline contains just isooctane and n-heptane in these
proportions, but that it has the same tendency to knock as this mixture. The blends
of known octane ratings can also be used as references as shown in Fig. 15.8. The
higher the number, the less likely is a fuel to pre-ignite.
15.4 Gasoline 311
Fig. 15.8 Octane rating (octane number) is determined by comparing compression ratio causing
knock using reference blends
ASTM D2699 and ASTM D2700 describe methods for measuring research octane
number (RON) at 600 rpm, which is most relevant to low speed city driving with load
conditions; and motor octane number (MON) at 900 rpm, which is most relevant to
high-speed highway driving conditions.
In North America, the posted octane of gasoline is the arithmetic average of RON
and MON: (R + M)/2. This is the number displayed on pumps at filling stations.
Typical grades are “regular” with a posted octane of 87, “mid-grade or medium”
312 15 Natural Gas and Petroleum Products
with a posted octane of 89, and “premium” with a posted octane of 91–93. In some
locales, customers can dial in any octane they want between 87 and 93.
The typical boiling range of gasoline is from 100 °F (~40 °C) to 400 °F (~200 °C).
The components have carbon numbers ranging from C5 to C12 . Gasoline is typ-
ically blended from several refinery streams including straight-run light naphtha
and products from several upgrading processes – isomerization (isomerate), alky-
lation (alkylate), catalytic reforming (reformate), polymerization (poly gasoline or
polymerate), catalytic cracking (cat gasoline), hydrocracking (hydrocrackate) and
coking. At some point, every component of gasoline goes through a hydrotreater or
hydrocracker, and heavy naphtha must go through a catalytic reformer. Other streams
undergo other treatments before they can be blended.
A gasoline containing a high proportion of straight chain alkanes has a greater ten-
dency to knock. However, branched-chain alkanes, cycloalkanes and aromatic hydro-
carbons are much more resistant to knocking. Straight-chain alkanes are converted
into isoalkanes in several processes in the refinery, as mentioned before. The octane
rating of gasolines usually available for cars contain a mixture of straight-chain,
branched, cyclic paraffinic and aromatic hydrocarbons, produced by the processes
described below.
Another important property of gasoline is Reid vapor pressure (RVP) which is
the vapor pressure of a motor gasoline measured at 100 °F in a volume of air 4 times
the liquid volume. It indicates the ease of starting and vapor-lock tendency as well
as explosion and evaporation hazards.
n
Mt (RVP) = Mi (RVP)i
i=1
where
Mt Total moles of blended product,
(RVP)t Specification RVP for product, psi
Mi Moles of component i
(RVP)i RVP of component i, psi or kPa.
In the mid- to late-20th century, making gasoline was a relatively simple task.
If a mixture of components met specifications for volatility and octane, it could be
shipped to retail outlets and sold as-was. If the octane was low, the problem could be
fixed by adding a small amount of tetraethyl lead (TEL). Butanes could be added or
left out as needed to adjust volatility, especially during winter time or cold weather.
Now, due to environmental regulations on hydrocarbon emissions, refiners must also
meet restrictions on RVP, olefins, sulfur content, and oxygen content. The new RVP
15.4 Gasoline 313
restriction greatly decreases the amount of butanes that can be used when ethanol is
present. In some locales during some seasons, not even pentanes can be blended into
gasoline without excessive RVP. As a result of these restrictions, especially the limit
on RVP, the air is much more breathable in large American and European cities.
As mentioned above, several refinery streams have the right vapor pressure, boiling
range, sulfur content and octane to end up in the gasoline pool. Table 15.4 shows
properties for blend stocks from which gasoline might be made. Ethanol has very low
RVP, compared to gasoline components. However, by blending 10% ethanol with
gasoline, it raises the RVP of the hydrocarbon component due to the phobic nature
of gasoline to ethanol [6]. RVP restrictions range from 7.8 to 9.0 psi. Depending on
location, there can be a 1 psi allowance when the ethanol content is 9–10 vol% [7].
Gasoline used for piston-engine powered aircraft is called aviation gasoline, or
avgas, which has high octane number and low flash point to improve ignition charac-
teristics on propeller aircrafts. As mentioned in the jet-fuel section, avgas can mean
either Jet B or JP-4.
A formulated (or finished) gasoline contains base fuel and small amounts of addi-
tives. Typical additives include metal deactivators, corrosion inhibitors, antioxidants,
detergents, demulsifiers, anti-icing additives and oxygenates. Table 15.5 lists the
additives used to prepare finished gasoline. Additive packages vary from season-
to-season, region-to-region, and retailer-to-retailer. After-market additives contain
314 15 Natural Gas and Petroleum Products
similar types of ingredients, and usually are more concentrated. They are packaged
to be added by consumers to their own vehicles.
Metal deactivators are compounds that sequester (deactivate) metal salts that oth-
erwise accelerate the formation of gummy residues during storage. The formation of
these gums is accelerated by metal salts, such as copper salt. These metal impurities
might arise from the engine itself or as contaminants in the fuel. Antioxidants, such
as phenylenediamines and other amines in 5–100 ppm, are used to prevent the degra-
dation of gasoline. Detergents are used to reduce internal engine carbon buildups,
improve combustion, and to allow easier starting in cold climates. Typical detergents
include alkylamines and alkyl phosphates at the level of 50–100 ppm.
Starting in the 1920s, tetraethyl lead (TEL) was added to gasoline in order to raise
engine compression as an octane booster, which improved the fuel economy. How-
ever, TEL had damaging effects on catalytic converters and was the main cause of
spark plug fouling. There was also a concern about the lead in air which could cause
brain damage especially for the younger children to touch surfaces contaminated by
lead. Therefore, TEL needed to be phased out, so refiners had to find other ways to
provide octane for gasolines. Gasoline blending became more complex in the 1970s.
In the 1990, the Clean Air Act was amended by Congress, requiring the phase-
out of TEL. It empowered the Environmental Protection Agency (EPA) to impose
emissions limits on automobiles. The reformulated gasoline (RFG) program was
mandated. RFG is gasoline blended to burn more cleanly than conventional gasoline
and to reduce smog-forming and toxic pollutants in the air, with sufficient octane
ratings. Smog or ground level ozone threatens the health and is particularly dangerous
to children and individuals with respiratory problems.
15.4 Gasoline 315
RFG was implemented in several phases. The Phase I program started in 1995
and mandated RFG for 10 large metropolitan areas. Several other cities and four
entire States joined the program voluntarily. Initially, the oxygen for RFG could be
supplied as ethanol or C5 –C7 ethers. The ethers have excellent blending octanes and
low vapor pressures. But due to leaks from filling station storage tanks, methyl-t-
butyl ether (MTBE) was detected in ground water samples in New York City, Lake
Tahoe, and Santa Monica, California. In 1999, the Governor of California issued
an executive order requiring the phase-out of MTBE as a gasoline component. That
same year, the California Air Resources Board (CARB) adopted California Phase 3
RFG standards, which took effect in stages starting in 2002. The standards include
a ban on MTBE and a tighter cap on sulfur content—15 wppm maximum.
Now, ethanol is the only oxygenate allowed to be blended into RFG. Legally,
ethanol must be denatured to avoid being classed (and taxed) as a liquor. The most
common recognized denaturant is gasoline, at 2–5% by volume. At this point, Almost
all of the gasoline in the US is reformulated. However, the blending ethanol into
gasoline reduces the total heat content of the RFG respective to an equal volume
of conventional gasoline, thus, reduces the fuel efficiency. Butanols, particularly
isobutanol and biobutanol (developed by BP and DuPont), have higher heating value
and octane ratings. But, only ethanol is used in the U.S. by mandate.
Tier 1 reformulated gasoline regulations required a minimum amount (2%) of
chemically bound oxygen, imposed upper limits on benzene and Reid Vapor Pressure
(RVP), and ordered a 15% reduction in volatile organic compounds (VOC) and air
toxics. VOC react with atmospheric NOx to produce ground-level ozone. Air toxics
include 1,3-butadiene, acetaldehyde, benzene, and formaldehyde.
The regulations for Tier 2, which took force in January 2000, were based on
the EPA Complex Model, which estimates exhaust emissions for a region based on
geography, time of year, mix of vehicle types, and—most important to refiners—fuel
properties. As of 2006, the limit on sulfur in the gasoline produced by most refineries
in the U.S. was 30 wppm, a 90% reduction from 300 wppm.
In the United States, Tier 3 vehicle emission and fuel standards lowered the
allowed sulfur content of gasoline from 30 wppm to 10 wppm, beginning in 2017. The
present limit on gasoline sulfur is 10 wppm.
The RFG program, combined with other industrial and transportation controls
aimed at smog reduction, is contributing to the long-term downward trend in U.S.
smog levels. About 75 million people breathe cleaner air because of RFG [8].
Volatility Minimum percent vaporized at 100 and 150 °C are 46 and 75,
respectively.
Hydrocarbon types The maxima for olefins/aromatics/benzene are 18/35/1 vol%,
respectively.
Oxygen The maximum oxygen content is 3.7% m/m. Oxygenates can
include different amounts of methanol, ethanol (with stabiliz-
ing agents), isopropyl alcohol, butyl alcohols, ethers such at
methyl-t-butylether (MTBE), and other mono alcohols.
Sulfur 10 ppmw.
15.5 Naphtha
Naphtha boils in the same boiling range as gasoline, between 30 and 200 °C, having a
similar carbon number range between 5 and 12. Light naphtha is the fraction boiling
between 30 and 90 °C and consists of molecules with 5–6 carbon atoms. Some light
naphthas, such as those from hydrocracking units, can be used as-is for gasoline
blending. Others are sent to isomerization units. Heavy naphthas boil between 80
and 205 °C and consist of molecules with 6–12 carbons. Most heavy naphthas are
feeds for catalytic reforming.
Naphtha boiling range products, including reformates, isomerates, and alkylates,
are used in high octane gasoline. Straight-run naphtha is used as a diluent in the
bitumen mining industry, as feedstock for producing olefins via steam cracking, and
as solvents for paints (as diluents), dry-cleaning, cutback asphalt, industrial extraction
processes, and in the rubber industry.
Kerosene, jet fuel, and turbine fuel have similar boiling ranges. The key product
properties are flash point, freezing point, sulfur content, and smoke point. The flash
point is the lowest temperature at which a liquid gives off enough vapor to ignite
when an ignition source is present. The freezing point is especially important for
jet aircraft, which fly at high altitudes where the outside temperature is very low.
Sulfur content is a measure of corrosiveness. The measurement of smoke point goes
back to the days when the primary use for kerosene was to fuel lamps. To get more
light from a kerosene lamp, you could turn a little knob to adjust the height of the
wick. But if the flame got too high, it gave off smoke. Even today, per ASTM D1322,
smoke point is the maximum height of smokeless flame that can be achieved with
calibrated wick-fed lamp, using a wick “of woven solid circular cotton of ordinary
quality.” The smoke point of a test fuel is compared to reference blends. A standard
40%/60% (volume/volume) mixture of toluene with 2,2,4-trimethylpentane has a
15.6 Kerosene and Jet Fuel (Turbine Fuel) 317
smoke point of 14.7, while pure 2,2,4-trimethylpentane has a smoke point of 42.8.
Clearly, isoparaffins have better smoke points than aromatics.
Kerosene boils between 150 and 275 °C (or 350 and 550 °F), with a typical carbon
number range between 6 and 16. Prior to the invention of automobiles, kerosene for
lighting was the most marketable product of petroleum. Today, the main uses of
kerosene include burning in lamps and domestic heaters or furnaces, as a fuel or fuel
component for jet engines or rockets, and as a solvent for greases and insecticides.
A jet engine, shown in Fig. 15.9, sucks air in at the front with a fan. The pressure
of the air is raised by a compressor, which is powered by a turbine connected to the
shaft, to reach a high temperature for fuel to burn. The air/fuel mixture is introduced
into fuel burner to combust, and the exhaust from the combustion chamber pushes
through the fans of the turbine and exits through the nozzle at the back of the engine.
As the jets of gas shoot backward, the engine and the aircraft are thrust forward (per
Newton’s third law of motion). Turbine engines can operate with a wide range of
fuels. Those with higher flash points are less flammable and safer to transport and
handle.
The primary sources of jet fuels are hydrotreated straight-run kerosene from an
atmospheric distillation unit and hydrocracker products with the right volatility. Jet
fuels are used in military and civil jets, with two groups of grading. Jet fuels have two
types: naphtha and kerosene jet fuels. Naphtha-type jet fuels are mainly for military
usage. The military grades are JP followed by a number, where JP stands for jet
propellant. Civil grades start with “Jet” followed by a letter.
In military jet grades [10], JP-1 was specified in 1944. It was pure kerosene, having
a freezing point of −60 °C. JP-2 and JP-3 were developed during World War II; they
are obsolete today. JP-2 had higher freezing point and JP-3 had higher volatility than
JP-1. JP-4 is a wide-cut fuel with a carbon number range of C4 to C16 , for broader
availability. JP-5 is specially blended kerosene. JP-6 is a higher cut than JP-4 with
fewer impurities. JP-7 is used in supersonic jets requiring higher flash point. JP-8 is
kerosene modeled on the Jet A-1 fuels used in civilian aircraft.
318 15 Natural Gas and Petroleum Products
For civilian jet fuels [11], Jet A and Jet A-1 are kerosene-type jet fuels. The primary
physical difference between the two is freeze point. Jet A specification fuel has been
used in the U.S. since the 1950s. Today, the Jet A and Jet A-1 specifications are the
same as those published by the International Air Transport Association (IATA). The
freezing point of Jet A is <−40 °C versus Jet A-1 at <−47 °C. Both Jet A and Jet
A-1 have a flash points higher than 38 °C (100 °F) with autoignition temperatures of
210 °C (410 °F). Kerosene-type jet fuel has a carbon number distribution between 8
and 16, similar to JP-8.
Jet B, also called aviation gasoline, is a lower-boiling fuel used for its enhanced
cold-weather performance. Naphtha-type jet fuel, sometimes referred to as “wide-
cut” jet fuel, has a carbon number distribution between 5 and 15, similar to JP-4.
The boiling range specifications for commonly used jet fuels are shown in
Table 15.6, along with other specifications.
Jet fuel additives include antioxidants, which are usually based on alkylated phe-
nols to prevent gumming; antistatic agents such as dinonylnaphthylsulfonic acid to
dissipate static electricity and prevent sparking; corrosion inhibitors; fuel system de-
icing agents; biocides to remediate microbial growth in the fuel system; and metal
deactivators, such as N,N -disalicylidine-1,2-propane diamine (MDA), to remediate
deleterious effects of trace metals, such as copper even at part-per-billion can catalyze
fuel oxidation, on the thermal stability of the fuel.
15.7 Diesel
Diesel fuels are used in diesel or compression ignition engines. Adiabatic compres-
sion of a fuel/air mixture provides enough heat for ignition; no spark is needed.
Compared to spark ignition engines, a diesel engine is more cost effective because of
its greater energy efficiency, higher power output, and better fuel economy under all
loads. With previous fuels, diesel engines were noisier and emissions of particulates
and nitrogen oxides (NOx ) were considerably higher.
Before 1993, when the United States Environmental Protection Agency (EPA)
began regulating diesel fuel, on-road diesel contained as much as 5000 parts per
million (ppm) of sulfur. Long-haul truck companies disposed of used crank-case
oil, which contained sludge, phosphorous-based additives, and other bad actors, by
blending it into diesel at their private terminals. The first step toward cleaner diesel
set a cap of 500 ppm sulfur for on-road diesel. Starting in 2006, EPA began to phase-
in regulations to lower the allowed sulfur to 15 ppmw. This fuel became known as
ultra-low-sulfur diesel (ULSD). At about the same time, similar regulations were
imposed in the UK, Europe, Japan, and other developed countries. Now in Europe
and the United States, the limit on sulfur content is 10 ppmw. The much stricter
diesel fuel standards had a huge impact on cleaning up diesel exhaust. One reason
is: sulfates formed by the combustion of organic sulfur compounds comprised a
significant proportion of particulates. Another reason: lower sulfur emissions allowed
the practical use of particulate traps and catalytic converters in diesel vehicles.
There are three categories of diesel fuels or oils in use: (1) land transportation
diesel fuels (officially called ‘onroad diesel’) used in trucks, buses, trains or other
land transportation vehicles that require high variation of speed and load; (2) marine
diesel fuels oils, used in ships that have variable speed but relatively high and uniform
load; and (3) plant or industrial diesel fuels, used in electric power generation plants
that have low or medium speed with heavy load [13]. Onroad diesel fuels used to be
known as No.1 diesel (super diesel) and No. 2 diesel (containing cracked oils with a
wider boiling range).
Modern automotive diesel specifications resemble those for what used to be called
No. 1 diesel. Now, the European Committee for Standardization defines Euro V diesel
in European Standard DIN EN 590. The 2017 specifications for Euro V diesel are
presented below [14, 15]. Note that the sulfur content must be <10 wppm, the density
must be <845 kg/m3 , and the cetane number must be >51. In China, diesel quality is
governed by China’s State Council. China V, which is similar to Euro V, is mandated
for onroad use in major urban areas, must contain <10 wppm sulfur. China IV, which
is allowed in other areas, must contain <50 wppm sulfur [16].
320 15 Natural Gas and Petroleum Products
Euro V diesel—EN590
Property Units low limit Upper Test-method
limit
Cetane index 46 – ISO 4264
Cetane number 51 – ISO 5165
Density at 15 °C kg/m3 820 845 ISO 3675, 12185
Polycyclic aromatic % (m/m) – 11 ISO 12916
hydrocarbons
Sulphur content mg/kg – 10 ISO 20846, 20847, 20884
Flash point °C 55 – ISO 20846, 20884
Carbon residue (on % m/m – 0.3 ISO 2719
90–100% fraction)
Ash content % (m/m) – 0.01 ISO 10370
Water content mg/kg – 200 ISO 6245
Total contamination mg/kg – 24 ISO 12937
Copper strip corrosion (3 h Rating Class 1 Class 1 ISO 12662
at 50 °C)
Oxidation stability g/m3 – 25 ISO 2160
Lubricity. Wear scar μm – 460 ISO 12205
diameter at 60 °C
Viscosity at 40 °C mm2 /s 2 4.5 ISO 12156-1
Distillation recovered at % V/V 85 <65 ISO 3104
250 °C. 350 °C
95% (V/V) recovered at °C – 360 ISO 3405
Fatty acid methyl ester % (V/V) – 7 ISO 14078
content
Truckers use higher density diesel, which used to be called ‘No. 2 diesel,’ to carry
heavy loads for long distances at sustained speeds because it’s less volatile than
automotive diesel and provides greater fuel economy.
Full-range gas oils boil between 150 and 400 °C (or 335–750 °F), and include
molecules with a carbon number range typically between 8 and 21. The boiling
range of automotive diesel is narrower. The initial boiling point is set by the flash
point, which must be greater than 55 °C. The flash point corresponds to a TBP initial
boiling point of about 180 °C (356 °F). The upper distillation limit is determined
by the ASTM D86 95% boiling, for which the maximum is 360 °C (680 °F). This
corresponds to a TBP final boiling point of 370–380 °C, depending on the nature of
the blend components of the fuel.
The most common type of diesel fuel comes from petroleum middle distillate, but
alternatives that are not derived from petroleum, such as biodiesel, biomass-to-liquid
(BTL) or gas-to-liquid (GTL) diesel, are increasingly being developed and becoming
more popular in North America.
15.7 Diesel 321
Diesel engines use compression heat to ignite a fuel/air mixture in the combustion
chamber. A typical 2-stroke diesel engine is shown in Fig. 15.10. At bottom dead
center (B.D.C.) the air enters the combustion center above the piston while exhaust
moves out of the chamber through open valves. With the exhaust valves closed, the
air is compressed to the lowest volume and the highest temperature at top dead center
(T.D.C.). This is the moment the fuel is injected and ignited to produce power. The
cycle is repeated at the B.D.C.
Just as gasoline is rated by its octane number, the ignition quality (knocking ten-
dency) of diesel fuels is measured on a single-cylinder rating engine by matching
fuel performance with standard blends. In the historical standard (ASTM D613-
10), blends of n-hexadecane with α-methylnaphthalene, where the cetane number is
defined as 100 for n-hexadecane (cetane) and 0 for α-methylnaphthalene, that pro-
vides the specified standard of 13° (crankshaft angle) ignition delay at the identical
compression ratio to that of the fuel sample. For example, a diesel fuel with cetane
number of 55 matches the performance of a blend of 55% of n-hexadecane and 45%
of α-methylnaphthalene in the cetane engine.
322 15 Natural Gas and Petroleum Products
Other important diesel-fuel properties include flash point, cloud point, pour point,
kinematic viscosity, lubricity—and of course sulfur content. Cloud point and pour
point indicate the temperatures at which the fuel tends to thicken and then gel in cold
weather. In a diesel engine, viscosity not only measures the tendency of a fluid to flow
but also indicates how well a fuel atomizes in spray injectors. Lubricity measures
the fuel’s quality as a lubricant for the fuel system, where it reduces friction between
solid surfaces (piston and cylinder) in relative motion. It indicates how the engine will
perform when loaded. Table 15.7 lists cetane numbers for selected pure compounds.
As with octane, blended cetane numbers can differ significantly from those for pure
compounds.
The U.S. EPA now requires that all nonroad, locomotive, and marine (NRLM)
diesel fuel must be ULSD, and that all nonroad, locomotive, and marine (NRLM)
engines and equipment must use ULSD (with some exceptions for older locomotive
and marine engines) [19].
Before 2020, marine diesel oil could contain some heavy fuel oil and even waste
products such as used motor oil. But as of 2020, MARPOL regulations will require
sea-going marine diesel to contain less than 0.5 wt% sulfur [20].
Most engines show an increase in ignition delay when the cetane number is decreased
from around 50–40. Adding 0.5 vol% of cetane improvers, such as alkyl nitrates,
primary amyl nitrates, nitrites, or peroxides, will increase the cetane number by 10
units. The viscosity of the fuel is important because many injection systems rely on
the lubricity of the fuel for lubrication. However, fuel lubricity will become poor after
usage, so polymeric lubricity agent is needed to maintain the viscosity. Cold flow
improvers, or flow-enhancing additives, provide important cold weather properties
of the desirable diesel fraction alkanes for cetane have melting points above 0 °C.
Diesel additives may contain cetane number improver, a lubricity agent, deter-
gents, dispersants, metal deactivators, and more. Table 15.8 lists the common addi-
tives used in diesel fuel and the reasons they are used.
Both heating oil and fuel oil are liquid petroleum products used as fuels for furnaces
or boilers to heat buildings, or used in generators to produce power. Larger normal
alkanes in heavier distillates burn with a less smoky flame and have higher flash
points than gasoline and kerosene, making them desirable for home heating fuels.
No. 1 fuel oil is a volatile distillate similar to kerosene, but with higher pour points
and end points. It is intended for ease of vaporization in pot-type burners. It has a
carbon number range of 9–16. No. 2 fuel oil is similar to No. 1 but contains cracked
stock, having a carbon number range of 10–20. No. 3 fuel oil is for burners requiring
low viscosity heating oil, which is described with No. 2 in specifications. No. 4 fuel
oil is usually a light residual oil used in a furnace that can atomize the oil and is not
equipped with a preheater. It has a carbon number range of 12–70. No. 5 fuel oil has
higher viscosities than No. 4, requires preheating to 170–220 °F for atomizing and
handling, and has a carbon number range of 12–70. It is also known as Bunker B oil.
No. 6 fuel oil is a high viscosity residual oil that requires preheating to 220–260 °F
for storage, handling, and atomizing, having a carbon number range of 20–70. It is
specified by the U.S. Navy as Bunker C oil for ships [21].
The fuel oils used in marine diesel engines have different classifications [21].
Marine gas oil (MGO) is made from distillate only. Marine diesel oil (MDO) is a
blend of heavy gasoil that may contain very small amounts of refinery residue feed
stocks, but it needs not be heated for use in internal combustion engines. Intermediate
fuel oil (IFO) is a blend of gasoil and heavy fuel oil, with less gasoil than marine
diesel oil. Marine fuel oil (HFO) is pure or nearly pure residual oil, roughly equivalent
to No. 6 fuel oil (Bunker oil).
Major seaports, including those in the U.S. and the European Union, are imposing
tighter emissions limits on sea-going ships and the fuels they burn. In 2012, a global
treaty (MARPOL Annex VI) capped fuel sulfur content at 1% in coastal waters and
extended the coastal zone to 200 miles offshore around Canada and the United States.
The sulfur limit was decreased to 0.1% on January 1, 2015, requiring even tighter
limits on the sulfur content of marine fuels. The treaty also imposes limits on SOx and
NOx emissions, which may require the use of stack scrubbers. Meanwhile, changes
in vessels and the use of larger, more fuel-efficient vessels have decreased marine
fuel consumption. As mentioned, as of 2020, all marine fuel oil used for international
shipping must contain <0.5 wt% sulfur.
Figure 15.11 gives an overview of product blending of fuels, gasoline, jet fuel
(kerosene), diesel and fuel oil, from various refining processes. Lubricant oils are
not obtained from distillation followed by various thermal and catalytic processes,
326 15 Natural Gas and Petroleum Products
as fuels. They are manufactured through various extraction processes; hence, their
products are discussed separately.
• Naphtha comes from the following units: distillation, isomerization, alkylation,
polymerization (not shown), catalytic reforming, hydrocracking, catalytic cracking
and coking. Naphthas from different units have similar boiling ranges but may have
contain different components. Some naphtha molecules might end up in gasoline,
some might be sold as solvent, and others could go to thermal cracking in an olefins
plant.
• Kerosene and jet fuel come from the hydrocracker unit (HCU) and the kerosene
hydrotreater (KHT).
• Gas-oil-boiling-range distillates come from the FCC, the HCU, the gas oil
hydrotreater (GOHT), visbreaker and coker. FCC gas oil sometimes is used as
fuel oil cutter stock, but otherwise it does not meet specifications for diesel. It
must undergo further refining before it can be sold. Almost every unit makes off-
gas. In the past, offgas streams were collected into a common system, desulfurized
and used for fuel gas.
Finished products are comprised of intermediate streams from different units,
along with stabilizers, antioxidants, corrosion inhibitors and other additives. To be
products, the blends must meet specifications developed and published by inter-
15.9 Blending of Fuels 327
national organizations such American Society for Testing and Materials (ASTM),
the International Organization for Standardization (ISO), the China State Council,
or similar agencies in other countries. Specifications incorporate information from
suppliers, users, government agencies and equipment manufacturers.
A material in the kerosene boiling range is a jet fuel if it meets ASTM D1655—15
specifications no matter how it is made. One might assume that a material is gasoline
if it meets the specifications of ASTM D4814—14b, no matter how it is made. But
in the United States, the assumption is wrong. By law, U.S. gasoline must contain
ethanol.
Lubricants were introduced in Chap. 4. Refiners prepare lube base stocks from dif-
ferent oils by removing asphaltenes, aromatics, and waxes. Lube base stocks are
hydrofinished, blended with other distillate streams for viscosity adjustment, and
compounded with additives to produce finished lubricants. In the past, solvent-based
technology was used to prepare lube base stocks. Propane deasphalting was used
to remove asphaltenes. Furfural and related substances were used to extract aro-
matics, and MEK or MIBK were used to remove wax. With the advent of cat-
alytic dewaxing (CDW), some or all of these solvent-based methods can be replaced
with hydroprocessing. CDW was developed by Mobil (now part of ExxonMobil)
in the 1980s. The Mobil process employs ZSM-5, which selectively converts waxy
n-paraffins into lighter hydrocarbons. The Isodewaxing Process, commercialized in
1993 by Chevron, reduces wax catalytically by isomerising n-paraffins into isoparaf-
fins. Isodewaxing also removes sulfur and nitrogen and saturates aromatics. The
products have a high viscosity index (VI), low pour point, and excellent response to
additives.
Mineral oil-based lubricant base oils can be categorized into 3 groups [22, 23].
Group I is manufactured by solvent extraction, solvent or catalytic dewaxing, and
hydro-finishing processes. It contains <90% saturates and >0.03% sulfur with vis-
cosity index (VI) of 80–120. Common Group I base oils are 150 N (solvent neutral),
500 N, and 150BS (bright stock). Group II is manufactured by hydrocracking or
catalytic dewaxing processes. It contains over 90% saturates and under 0.03% sulfur
with VI of 80–130. Group II base oil has superior anti-oxidation properties since
virtually all hydrocarbon molecules are saturated. It has water-white color. Group
III is manufactured by special processes such as hydroisomerization and from base
oil or slack wax through dewaxing processes. It also contains >90% saturates and
<0.03% sulfur, but with VI >120.
328 15 Natural Gas and Petroleum Products
Synthetic lubricant base oils are also derived from petroleum with VI >120. In
North America, Groups III, IV and V are now described as synthetic lubricants, with
group III frequently described as synthesized hydrocarbons, or SHCs. In Europe,
only Groups IV and V may be classed as synthetics. Group IV is specific to the
base oils derived from poly alpha olefins (PAOs). Group V includes other synthetic
oxygen-containing base stock, such as dibasic esters, aromatic esters, polyol esters,
etc.
The lubricant industry commonly extends this group terminology to include:
Group I+ with a VI of 103–108, Group II+ with a VI of 111–119, Group III+
with a VI of at least 130, and Group IV+ with VI 5–15 higher than conventional
PAOs made strictly from 1-decene [24].
Finished lubricants are formulated with base oil and special additives. Detergents
and dispersants are the dominant performance additives in passenger car motor
engine oils. The additives are typically 5–20% of formulated engine oil or 55–70%
of performance-package oil. The remaining additives include antiwear agents, ash
inhibitors, friction modifiers, etc. [25].
Finished lubricant oils with additives can be used as engine oils, transmission
fluids, gear oils, turbine oils, hydraulic oils, metal cutting oils, and paper machine
oils. They can be mixed with soaps to make greases.
Lubricating oil additives are used to enhance the performance of lubricants and func-
tional fluids. Each additive is selected for its ability to perform one or more specific
functions in combination with other additives. Selected additives are formulated into
packages for use with a specific lubricant base stock and for a specified end-use
application. The largest end use is in automotive engine crankcase lubricants. Other
automotive applications include hydraulic fluids and gear oils. In addition, many
industrial lubricants and metalworking oils also contain lubricating oil additives.
Additives are organic or inorganic compounds dissolved or suspended as solids
in oil. They typically range between 0.1 to 30% of the oil volume, depending on the
machine. The formulated lubricant oils in the automotive engine crankcase comprise
of up to 5–10% of additives. Oil additives are vital for the proper lubrication and pro-
longed use of motor oil in modern combustion engines. Some of the most important
additives include those used for viscosity and lubricity, for the control of chemical
breakdown, and for seal conditioning. Some additives permit lubricants to perform
15.10 Lubricant Base Oils, Wax, Grease and Specialty Products 329
better under severe conditions, such as extreme pressures and temperatures and high
levels of contamination.
The major functional additive types in the additive package are dispersants
(40–50%), detergents (15–20%), oxidation inhibitors (antioxidants), antiwear agents,
ashless inhibitors, friction modifiers, corrosion inhibitors, antifoam agents, demulsi-
fying agents, pour point depressants, metal deactivators, extreme-pressure additives,
tackiness agents, and viscosity index improvers. The dispersant and detergent make
up about 50–70% of the package. Hence, the chemistry of the total package and the
finished formulated oils is greatly influenced by these components.
It should be noted that it is not always better to use more additives. As more additive
is blended into the oil, sometimes there isn’t any more benefit gained, and at times
the performance of the blend actually deteriorates. In other cases, the performance
of the additive doesn’t improve, but the duration of service does improve.
In addition, increasing the percentage of a certain additive may improve one prop-
erty of an oil while at the same time degrade another. When the specified concentra-
tions of additives become unbalanced, overall oil quality can also be affected. Some
additives compete with each other for the same space on a metal surface. If a high
concentration of an anti-wear agent is added to the oil, the corrosion inhibitor may
become less effective. The result may be an increase in corrosion-related problems.
Some of the additives are discussed below.
15.10.3.1 Detergents
Corrosion inhibitors protect the metal surface against corrosion in a variety of lubri-
cant applications. Some inhibitors neutralize acids; others form protective films.
Detergents are excellent corrosion inhibitors because they protect in both ways.
Some detergents can also act as oxidation inhibitors.
The functions of detergents are to solubilize polar components, inhibit corrosion,
and prevent high temperature deposits, in part through neutralization of acids. They
are composed of two components: surfactants (organic), generally sulfonates, phen-
ates or salicylates; and a colloidal metal phase (inorganic), generally Ca, Mg or Na
from the over-basing, where the level of the basic phase is high relative to the amount
of surfactant.
The combination of a surfactant molecule with a colloidal inorganic core results in
a micellular-type structure shown in Fig. 15.12, with the polar head of the molecule
represented by a circle. The long nonpolar alkyl tail (C12 –C32 ) provides good solubil-
ity of nonpolar base stock molecules. The amorphous colloidal phase is represented
by CaCO3 in the figure. This basic colloidal carbonate neutralizes the acids formed
during combustion. The inorganic acids are nitric acid, sulfuric acid and hydrochloric
acid, which lead to metal corrosion and wear. Organic acids lead to polymerization,
viscosity increase and resin formation. The detergent adheres to dirt and oil insoluble
products formed as oxidation by-products during equipment operation. They keep
these in suspension, preventing them from depositing onto critical engine surfaces.
330 15 Natural Gas and Petroleum Products
Together, the surfactant and the basic components form a protective layer that
inhibits corrosion, inhibits oil degradation, reduces high temperature deposits, and
solubilizes polar contaminants.
Over-based detergents neutralize acidic combustion products, helping to control
corrosion and resinous build-up in the engine. However, the total base number will
continue decreasing during the use of oils, leading to an increase in total acid number
in the oil. The equivalence point of total base number and total acid number occurs
at around 3000–6000 miles of driving. At this point, the acids build up reaches
unacceptable levels. It is therefore desirable to change the oil before the total base
number and total acid number cross.
15.10.3.2 Dispersants
Extreme pressure (EP) additives are sulfur-phosphorus based chemicals that form
organo-metallic salts on the loaded surfaces during operation at high pressures and
temperatures. They create a protective layer that reduces wear between two mating
332 15 Natural Gas and Petroleum Products
O
PIB
H H
N N
N N NH 2
H
A dispersant molecule: polyisobutylene (PIB) maleic amide with tetraethylene pentaamine (TEPA)
metal surfaces. Anti-wear additives perform in a similar manner, but tend to operate
under lower pressures and temperatures. EP additives are usually supplemented with
anti-wear additives to make the oils, especially gear oils and grease. They are effective
across a wide range of pressure and temperature conditions. Friction modifiers are
mild anti-wear agents that minimize light surface contacts (sliding and rolling). They
prevent scoring and reduce wear and noise by forming protective surface layers. They
can be esters, natural and synthetic fatty acids as well as some solid materials such as
graphite and molybdenum disulfide. The organic molecules have a polar end (head)
and an oil-soluble end (tail). Once placed into service, the polar end of the molecule
finds a metal surface and attaches itself. These are also called boundary lubrication
additives. The inorganic molecules (graphite and MoS2 ) are highly stable and are
comprised of flat layers that slip easily over each other.
The most important advance in antiwear chemistry was made during the 1930s
and 1940s with the discovery of zinc dialkyldithiophosphates (ZDDP). These com-
pounds were found to have exceptional antioxidant and antiwear properties. The
antioxidant mechanism of the ZDDP was the key to its ability to reduce bearing
corrosion as a corrosion inhibitor. Since the ZDDP suppresses the formation of per-
oxides, it prevents the corrosion of Cu/Pb bearings by organic acids. Antiwear and
extreme-pressure additives function by thermally decomposing to yield compounds
that react with the metal surface. These surface-active compounds form a thin layer
that preferentially shears under boundary lubrication conditions.
15.10 Lubricant Base Oils, Wax, Grease and Specialty Products 333
The rate of oxidation doubles with every 10 °C increase in temperature. In the case
of increased temperature, in the presence of oxygen and mechanical load, lubricants
age very readily. If not controlled, the radical-induced oxidations generate acids,
which lead to the molecular degradation and finally to the failure of the lubricant.
The lubricant decomposition will lead to oil thickening and the formation of sludge,
varnish, resin and corrosive acids. This leads to an increase in corrosion. Water and
polar impurities increase the speed of attack, and internal combustion engines con-
tain plenty of these contaminants. Incorporating an oxidation inhibitor will interrupt
and terminate the free radical process of oxidation and thus slow down the age-
ing of lubricants. Phenolic materials are quite good for this purpose and the two
most commonly used inhibitors are 2,6-ditertiary-butylphenol (DBP) and 2,6-di-
tertiary-butyl-4-methylphenol or 2,6-di-tertiary-butyl-paracresol (DBPC) which is
also known as butylated hydroxytoluene (BHT). The typical recommended value of
DBPC and DBP in fresh oil is approximately 0.3% by weight.
Mineral oil lubricants become less effective at high temperatures because heat
reduces their viscosity and film-forming ability. Viscosity index improvers and thick-
eners are polymeric molecules, such as olefin copolymers (OCP), and are added to
reduce lubricant viscosity changes at high and low temperatures. OCP are copolymers
of ethylene and propylene. When polymer coils interact with oil and each other they
become increasingly resistant to flow; hence, they can be added to oils to increase
their viscosity (thickening). A balance between the thickening efficiency and shear
stability of the polymer is important. Higher molecular weight polymers make better
thickeners, but tend to have less resistance to mechanical shear. Lower molecular
weight polymers are more shear resistant, but do not improve viscosity as effectively
at higher temperatures and must be used in larger quantities. Mineral oil lubricants
become less effective at high temperatures because heat reduces their viscosity and
film-forming ability.
When viscosity index improvers are added to low-viscosity oils, they effectively
thicken the oil as temperature increases. This means the lubricating effect of mineral
334 15 Natural Gas and Petroleum Products
oils can be extended across a wider temperature range. At low temperatures, the
molecule chain contracts and does not impact the fluid viscosity. At high temper-
atures, the chain relaxes and an increase in viscosity occurs. Viscosity improvers
are primarily used in multigrade engine oils, gear oils, automatic transmission flu-
ids, power steering fluids, greases and various hydraulic fluids. Most of these uses
involve automobiles in which they are subjected to tremendous temperature swings
over short periods.
Polymer additives can undergo thermal and oxidative degradation, unzipping back
to smaller monomers, which reduces their effect. Hence, the highest possible degree
of thermal and oxidative stability is desirable.
Most lubricant applications involve agitation, which traps air in the lubricant and
encourages the formation of foam. Surface-active materials, such as dispersants
and detergents, further increase foaming tendency. Excessive foaming may result
in increased lubricant oxidation and decreased operational efficiency.
Antifoaming agents, such as poly dimethyl siloxane and copolymers of polyethy-
lene glycol and polypropylene glycol, alter the surface tension of the oil and help
to weaken the structure of air bubbles. The result is better lubricating qualities and
reduced maintenance.
15.10 Lubricant Base Oils, Wax, Grease and Specialty Products 335
Although most of the wax is removed during base oil dewaxing, high-molecular
weight species are necessary to achieve the desired target viscosity. A range of pour
point depressants based on polymethacrylate (PMA) and styrene ester in lubricant
formulation prevent wax fractions in the base oil from forming large crystal networks
which inhibit lubricant flow at cold temperatures, while keeping the viscosity benefits
at higher temperatures.
15.10.3.10 Demulsifiers
There are great improvements in lubricant base oils. However, a good additive pack-
age is very important for high-quality performance. To design a lubricant additive
package, we need to consider the major interaction between base oil and additives,
which is shown in Fig. 15.15. It’s not a complete picture, but at least it shows the
336 15 Natural Gas and Petroleum Products
major additive functions. Compatibility of additives to base oil and with other addi-
tives must be considered.
15.10.5 Greases
Greases are made by blending detergents (salts of long-chained fatty acids) into
lubricating oils at 400–600 °F (204–315 °C). Antioxidants are added to provide
stability. Some greases are batch-produced, while others are made continuously. The
characteristics of a grease depend to a great extent on the counter-ion (calcium,
sodium, aluminium, lithium, etc.) in the fatty-acid salt.
Other specialty products include white oil, insulating (electrical) oils and insecticide.
Naphthenic crude oils give white oil products of high specific gravity and viscosity,
suitable for pharmaceutical uses. Paraffinic crude oils produce oils with lighter grav-
ity and lower viscosity suitable for lubrication purposes. Technical grade white oils
are employed for cosmetics, textile lubrication, insecticides, vehicles, paper impreg-
nation, etc. Pharmaceutical (medical) grade white oils may be employed as laxatives
or for lubrication of food-handling machinery. Insulating oils are highly refined
fractions of low viscosity and are stable at high temperature. They have excellent
electrical insulating properties, so they can be used in transformers, circuit breakers,
15.10 Lubricant Base Oils, Wax, Grease and Specialty Products 337
and oil-filled cables, and for impregnating the paper covering of wrapped cables.
Insecticides are petroleum oils in water-emulsion form to be sprayed onto fruit trees
and swamp water to killing certain species of insects.
Wax recovery processes were introduced in Chap. 13. Gas oils or vacuum distillates
are the source of paraffin waxes, which forms large crystals or plates with molecular
weight between 300 and 450 (C17 –C35 ). Such wax contains mostly normal paraffins
and small amounts of isoparaffins, cycloparaffins and minute amounts of aromat-
ics. Vacuum residue is the source of microcrystalline wax with molecular weights
between 450 and 800 (C35 –C60 ). Microcrystalline wax contains mostly cycloparaf-
fins with n-alkyl and isoalkyl side chains. Paraffin wax is used for candles, cosmetics,
and other purposes. Microcrystalline waxes are used in the tire and rubber indus-
tries, where they are sometimes blended with paraffin wax. Food-grade wax is a
highly refined product, which is employed for water-proofing beverage containers,
for orthodontics and pharmaceuticals, and as food additive.
15.12 Cokes
Green coke is a raw carbonaceous solid of petroleum coke (petcoke) derived from
delayed coking. Green coke is calcined to remove volatile hydrocarbons and sulfur
compounds. With sufficiently low metal content it can be calcined for use as anode
materials (anode grade coke). Green coke with a too-high metal content is used as
fuel (fuel grade coke).
Undesirable coke is deposited on catalysts used in certain oil refining processes,
such catalytic reforming, hydrotreating and hydrocracking. In the FCC process, coke
is not undesirable; it is essential, because burning the coke provides heat to run the
unit.
The most common grades of finished coke are fuel grade coke, calcined petroleum
coke, and needle coke. Fuel coke comes from delayed coking units in the form of
sponge coke, which is named for its sponge-like appearance. It is produced from
coker feeds with low-to-moderate asphaltene concentrations. A good fuel coke with
high heat value and low ash content is used for power generation. Low ash content is
important, because ash can foul boilers. (The low ash content of fuel coke makes it
superior to coal for power generation.). When burning fuel coke, some form of sulfur
capture is required to meet current North American emission standards. Fuel grade
coke can be a feedstock for gasification, during which it is converted into synthesis
gas—a mixture of CO and H2 .
Sponge coke is the most common petroleum coke. In its uncrushed state, sponge
coke closely resembles coal. If sponge coke meets certain specifications, such as low
338 15 Natural Gas and Petroleum Products
metals content, it may be calcined and formed into carbon anodes for the aluminum
industry. Otherwise, it serves as a fuel. Sulfur recovery is required during calcination,
too.
Needle coke, named for its needle-like crystalline structure, is a high-value product
made from feeds that contain nil asphaltenes, such as hydrotreated FCC decant
oils. Needle coke has a low thermal expansion coefficient, making it suitable for
conversion into graphite electrodes for the electric-arc furnaces used in the steel
industry. It is less typically to break than the other types of petroleum cokes.
Honeycomb coke is an intermediate between sponge coke and needle coke. It
is characterized by ellipsoidal pores uniformly distributed throughout its shape. It
has lower electrical conductivity and a lower thermal coefficient compared to needle
coke.
Shot coke is undesirable because it tends to be unstable. It forms when the con-
centration of feedstock asphaltenes and/or coke-drum temperatures are too high. A
block of shot coke is a cluster of discrete mini-balls 0.1–0.2 in (2–5 mm) in diameter.
The clusters can be as large as 10 in (25 cm) across. If a cluster breaks apart when
the coke drum is opened, it can spray a volley of hot mini-balls in every direction.
The name “shot” derives from the fact that it can resemble shotgun ammunition.
Adding aromatic feeds, such as FCC decant oil, can eliminate shot coke formation.
Other methods of eliminating shot coke—decreasing temperature, increasing drum
pressure, increasing the amount of product recycle—decrease liquid yields, which
is not desired.
Specialty carbon products made from petroleum include recarburizer coke, which
is used to make special steels, and titanium dioxide coke, which is used as a reducing
agent in the titanium dioxide pigment industry.
Because of the similar coal-like properties, the demand of petroleum coke is rising
in China, Mexico, Canada and Turkey. The U.S. is becoming a large net exporter of
petroleum coke.
15.13 Asphalt
Asphalt is a sticky, black and highly viscous liquid or semi-solid form of petroleum.
It may be found in natural deposits or may be a refined product. Humans used asphalt
to make spears and coat baskets as far back as 70,000 years ago near Umm el Tlel, in
present-day Syria. As early as 3000–4000 BC, Egyptians used pitch to grease chariot
wheels and asphalt to embalm mummies. Mesopotamians used bitumen to line water
canals, seal boats, and build roads. Today, the primary use of asphalt/bitumen is
in road construction (paving), where it serves as the glue or binder that is mixed
with aggregate particles (stone, sand, polymer, etc.) to create asphalt concrete. Other
main uses are for waterproof roofing products.
Asphalt base stocks can be produced directly from vacuum residue or by solvent
deasphalting. Vacuum residue is used to make road-tar asphalt. To drive off light
15.13 Asphalt 339
ends, it is heated to about 750 °F (400 °C) and charged to a column where a vacuum
is applied to prevent thermal cracking.
In road-paving, the petroleum residue serves as a binder for aggregate, which
can include stone, sand, or gravel. The aggregate comprises about 95% of the final
mixture. Polymers are added to the binder to improve strength and durability. The
recommended material for paving highways in the United States is Superpave 32 hot-
mix asphalt. Superpave was developed in 1987–93 during a US$50 million research
project sponsored by the Federal Highway Administration.
Roofing asphalt is produced by bubbling air through liquid asphalt at 260 °C
(500 °F) for 1–10 h. During this “blowing” process, organic sulfur is converted to
SO2 . Catalytic salts such as ferric chloride (FeCl3 ) may be added to adjust product
properties and increase the rates of the blowing reactions, which are exothermic. To
provide cooling, water is sprayed into the top of the blowing vessel, creating a blanket
of steam that captures sulfur-containing gases, light hydrocarbons, and other gaseous
contaminants. These are recovered downstream. Cooling water may also be sprayed
on the outside of the vessel. The length of the blow depends on desired product
properties, such as softening temperature and penetration rate. A typical plant blows
four to six batches of asphalt per 24-h day. There are two primary substrates for
roofing asphalt—organic (paper felt) and fiberglass. The production of felt-based
roofing shingles consists of:
• Saturating the paper felt with asphalt
• Coating the saturated felt with filled asphalt
• Pressing granules of sand, talc or mica into the coating
• Cooling with water, drying, cutting and trimming, and packaging.
If fiberglass is used as the base instead of paper felt, the saturation step is elimi-
nated.
15.14 Petrochemicals
An overview of refinery processes for making products, including gases, fuels, lubri-
cants, grease, wax, coke and asphalt is shown in Fig. 15.16.
Petroleum products by estimated carbon number ranges are shown in Fig. 15.17.
Note that the upper carbon number range is not certain due to the limits of current
analytical instruments. The use of high temperature gas chromatography (HT-GC)
and field ionization/field desorption mass spectrometry (FI/FD MS), for example, can
determine the upper carbon numbers for lubricant oils and asphalts beyond the figure
shown. In addition, the ranges shown are only used as reference, which can vary at
different refineries and countries to meet products specifications or regulations.
Fig. 15.17 Petroleum products by estimated carbon number with uncertainty in heavy products
[28]
References 341
References
24. Wu MM, Ho SC, Luo S (2017) Synthetic lubricant base stock. In: Hsu CS, Robinson PR (eds)
Springer handbook of petroleum technology. Springer, New York (Chapter 35)
25. Burrington JD, Pudelski JK, Roski JP (2006) Challenges in detergents and dispersants for
engine oils. In: Hsu CS, Robinson PR (eds) Practical advances in petroleum processing.
Springer, New York (Chapter 18)
26. Petrowiki: oil demulsification. http://petrowiki.org/Oil_demulsification. Accessed 28 Nov 2016
27. Al-Salem SM, Ma X, Al-Mujaibel MM (2017) Carbon dioxide mitigation. In: Hsu CS, Robin-
son PR (eds) Springer handbook of petroleum technology. Springer, New York (Chapter 32)
28. ALS environmental: petroleum hydrocarbon ranges. http://www.caslab.com/Forms-
Downloads/Flyers/PETROLEUM_HYDROCARBON_RANGES_FLYER.pdf. Accessed
23 Nov 2016
Part IV
Petrochemicals, Midstream,
Safety and Environment
Chapter 16
Petrochemicals
The chemicals that can be produced from methane are summarized in Fig. 16.2.
CH4 + H2 O ↔ CO + 3H2
Ammonium
Nitrate
Carbon
Pyrolysis Black
Chlorine
Chloroethylene Methyl
Alkaline Chloride
Vinyl Chloride/
HCl or Acetic Acid Vinyl Acetate Methylene
Dichloride
Vinyl HCl
Chloroprene
Acetylene Chloroform
Cl2
Carbon
Tetrachloride
CO + H2 O ↔ H2 + CO2
The first reaction is highly endothermic, due to the high thermochemical stability
of methane and water. Under the conditions shown, the LTS reaction is slightly
exothermic.
The carbon dioxide is then removed either by absorption in aqueous ethanolamine
solutions, by adsorption into molten potassium carbonate (Benfield process [2]), or
by a pressure swing adsorption (PSA) unit. PSA hydrogen usually is 99.99% pure.
For many catalytic processes, CO is a catalyst poison and must be removed.
If required, catalytic methanation removes any small residual amounts of carbon
monoxide or carbon dioxide from the hydrogen:
Methanol also is made from synthesis gas (syngas), a mixture of CO, H2 and very
often some CO2 , that can be produced by partial oxidation of methane, coal, vacuum
residue, and biomass.
The methanol thus formed may be converted to gasoline by the Mobil (ExxonMo-
bil) process developed in the early 1970s [3]. First methanol is dehydrated to give
dimethyl ether (DME):
348 16 Petrochemicals
The ether is then dehydrated along with methanol over a zeolite catalyst, ZSM-5,
to yield an organic product containing up to 80 wt% C5 + hydrocarbons (paraffins,
naphthenes and aromatics). Reactions include polymerization and hydrogenation of
the ethylene intermediate. Stabilized gasoline, which produced by removing flue gas
and LPG, is split into light and heavy fractions. The heavy gasoline is hydrotreated
to reduce durene (1,2,4,5-tetramethylbenzene) and then blended back with the light
gasoline. ZSM-5 is deactivated by a carbon build-up (“coking”) over time. The
catalyst can be re-activated by burning off the coke in a stream of hot air at 500 °C;
however, the number of re-activation cycles is limited.
The processes of making gasoline from methane or syngas through methanol by
the MTG process and making diesel and jet fuels from syngas by the Fischer-Tropsch
process are also known as gas-to-liquids (GTL) processes [4].
At lower temperatures, methanol reacts to form DME. At higher temperatures,
olefins are produced and the selectivity for DME decreases. The zeolite catalyst
for methanol-to-olefins (MTO) is typically H-SAPO-34. In SAPO-34, some silicon
atoms occupying sites in ZSM-5 are replaced by phosphorus atoms so that aluminum
and phosphorus occupy the sites alternatively. The pore size is about 3.8 Å, compared
to 5.5 Å for ZSM-5. SAPO-34 also has weaker acidity than ZSM-5. Hence, ZSM-5
is used for MTG and SAPO-34 is used for MTO to yield polymer grade ethylene and
propylene for the production of polyolefins.
The first step in ammonia synthesis is to produce hydrogen, usually from natural
gas (methane) in SMR as described above. If present, H2 S is first removed from
natural gas by amine adsorption. Then, organic sulfur compounds undergo catalytic
hydrogenation, which converts them to gaseous hydrogen sulfide:
H2 + RSH → RH + H2 S (gas)
The gaseous hydrogen sulfide is then absorbed and removed by passing it through
beds of zinc oxide, where it is converted to solid zinc sulfide:
H2 S + ZnO → ZnS + H2 O
3H2 + N2 → 2NH3
Ammonia can be transported as anhydrous liquid ammonia, which boils at −33 °C.
Hence, it must be kept cool and under modest pressure. Ammonia is used primarily
for fertilizer. Anhydrous ammonia can be applied directly to soil. Alternatively,
ammonia can be converted to ammonium nitrate (NH4 NO3 ) or urea (NH2 –O–NH2 )
and blended with other compounds to make solid or aqueous fertilizers with defined
ratios of nitrogen, potassium, and phosphorous. Ammonia is not only one of the
350 16 Petrochemicals
main components of fertilizers, but it also is widely used in other sectors, such as the
pharmaceutical, household cleaning, and fermentation industries.
4NH3 (g) + 5O2 (g) → 4NO (g) + 6H2 O (g) (H = −905.2 kJ)
Nitric oxide is then reacted with oxygen in air to form nitrogen dioxide.
This is subsequently absorbed in water to form nitric acid and nitric oxide.
The main industrial use of nitric acid is for the production of fertilizers. The other
main applications are for the production of explosives, nylon precursors, and spe-
cialty organic compounds. Nitric acid is neutralized with ammonia to give ammo-
nium nitrate, a fertilizer. Ammonium nitrate is explosive. It has been responsible
for horrendous industrial accidents. When mixed with fuel oil, it becomes ANFO
(ammonium-nitrate-fuel-oil), the material used to attack the Murrah Federal Building
in Oklahoma City on April 19, 1995 [10].
The production of urea, also known as carbamide, from ammonia is carried out by
two steps. In the first step, liquid ammonia reacts with dry ice to form ammonium
carbamate (H2 N–COONH4 ) in an exothermic reaction:
In the second step, ammonium carbamate is decomposed into urea and water, an
endothermic reaction:
16.1 Chemicals from Methane 351
H2 N−COONH4 ↔ (NH2 )2 CO + H2 O
Vinyl chloride can be produced from acetylene and hydrogen chloride using mercuric
chloride as a catalyst. This method has been superseded by the more economical
processes based on ethylene in the United States and Europe. It remains the main
production method in China. Vinyl chloride is a toxic and carcinogenic gas that has
to be handled with special protective procedures.
Polyvinyl chlorides (PVC’s) are the third-most widely produced plastics, after
polyethylenes (PE) and polypropylenes (PP). PVC is made from vinyl chloride
(chloroethane), a product of the reaction between ethylene with oxygen and hydrogen
chloride over a copper catalyst.
PVC is used as a corrosion-resistant material for pipes in modern plumbing, siding
for houses, decking, and many other applications in construction material supply
chains. Adding plasticizers, mostly phthalates, make them flexible for use in clothing,
upholstery, electric cable insulation, inflatable products, and rubber replacement.
Flexible PVC is also used in medical tubing to replace glass and rubber. PVC is
352 16 Petrochemicals
better because it is more resistant to other chemicals than rubber and, unlike glass,
it does not break.
The precursor of polyvinyl acetate, vinyl acetate, can also be produced from acety-
lene via the gas-phase addition of acetic acid in the presence of metal catalysts, such
as mercury(II) catalysts. Like vinyl chloride, the industrial route has been superseded
by the ethylene-based process which will be discussed later.
CH3 OH → CH2 O + H2
The most important process for producing hydrogen cyanide is the reaction of
methane and ammonia in the presence of oxygen at about 1200 °C over a Pt catalyst
(Andrussow oxidation [14]):
Hydrogen cyanide (HCN) gas or vapor is highly toxic and fatal in sufficient con-
centrations. It was used in chemical warfare and is now prohibited for that purpose by
international laws. It is a highly valuable precursor to many chemical compounds
ranging from polymers to pharmaceuticals. Aqueous sodium cyanide, in concentra-
tions ranging from 100 to 500 parts per million, is used in gold and silver mining.
Cyanide salts also are used in electroplating these metals. HCN reacts with aldehy-
des and ketones to form cyanohydrins (R2 C(OH)CN), which are intermediates in
many organic syntheses. It reacts with ethylene oxide to form an intermediate which
eventually is converted to acrylonitrile.
The major component of natural gas is methane, accompanied by natural gas liquids
(NGL), i.e., ethane, propane, butanes and natural gasoline (C5 +). The components
of NGL are separated in a gas processing plant through absorption, condensation,
fractionation, or other methods. The separation of ethane, propane, and butanes is
accomplished in a series of tall fractionation columns, similar to those shown in
Fig. 11.4.
Once the ethane is separated, it is shipped by pipeline to a cracker facility, which
is a very sophisticated series of processes that convert the ethane to ethylene. The
first step entails using steam to transport ethane or a mixture of ethane with a small
amount of propane to a series of cracking furnaces, where it is heated to approximately
1500 °F. This requires a lot of energy. At that temperature, dissociation of C–H or
C–C bonds generates free radicals, which through a chain of reactions are converted
to ethylene, hydrogen, and heavier products (Fig. 16.3). In commercial units, process
conditions are selected to minimize follow-on reactions, such as radical addition to
generate C3 + compounds and cyclization to generate aromatics, polyaromatics, and
coke. Typically, ethylene comprises about 80% of the total product. In the same
fashion, propane is converted to propylene and other products.
Alternative feeds to cracker furnaces include sulfur-free naphtha and highly
upgraded heavy oils, such as hydrocracker unconverted oil. Ethylene and other
olefins also come from refinery cracking units. Especially important are propylene
and butylenes from FCC units [15].
Due to its highly reactive double bond, ethylene is particularly well-suited for
many different chemical reactions. It is one of the most important chemicals in all
chemistry. The chemicals derived from ethylene are summarized in Fig. 16.4.
Ethanol can be produced by hydration of ethylene with high pressure steam at
300 °C (572 °F) in a 1.0:0.6 ratio of ethylene:steam, catalyzed by phosphoric acid
adsorbed onto a porous support such as silica gel or diatomaceous earth. Ethanol
can also be produced by reacting ethylene with sulfur acid to form ethyl sulfate,
which in turn reacts with water to form ethanol. Upon partial oxidation over a silver
catalyst at about 500–650 °C, ethanol can be converted to acetaldehyde. However, the
main production method is the oxidation of ethylene via the Wacker process which
involves oxidation of ethylene over a homogeneous palladium/copper catalyst:
Ethylene also reacts with other hydrocarbons in a chemical process called alky-
lation. Benzene reacts with ethylene over a highly acidic zeolite catalyst to form
ethylbenzene. Ethylbenzene is a derivative petrochemical building block found in
the manufacturing supply chains for styrofoam and automobile tires.
16.2 Ethane and Ethylene 355
Radical formation
CH3CH3 → 2 CH3•
Hydrogen abstraction
CH3• + CH3CH3 → CH4 + CH3CH2•
Radical decomposition
CH3CH2• → CH2=CH2 + H•
Radical addition
CH3CH2• + CH2=CH2 → CH3CH2CH2CH2•
Cyclization
CH3CH=CHCH2CH2CH2• → C6H6 (benzene) + 2 H2 + H•
H• + H• → H2
16.2.1 Polyethylenes
After purification, ethylene goes from the cracker facility to a nearby complex in
which it undergoes polymerization to polyethylene.
The polymerization units are custom-designed and operated to control the specific
physical properties of the products, which typically are formed into small plastic
pellets called polyethylene resins. The different resins can be transformed into diverse
products with many shapes, sizes, and colors. Products range from grocery bags,
plastic bottles, and plastic toys to sophisticated, high tech military helmets.
In addition to conventional plastics, polymers can be thermoplastics, thermosets,
elastomers, rubbers, and fibers. Thermoplastics can be shaped and molded easily
when heated, and they melt when they are hot enough. Thermoplastics possess a
glass transition temperature, Tg. They are soft and pliable above the Tg, and hard
and brittle below. The Tg is different for each type of polyethylene. Depending on
the crystallinity and molecular weight of a plastic, a melting point and Tg may or
may not be observable. Sometimes, additives called plasticizers are introduced to
make a plastic softer and more pliable.
356 16 Petrochemicals
LPDE are made by free radical vinyl polymerization, while HDPE are made cat-
alytically over alkyl aluminum and titanium (III) chloride in the Ziegler-Natta process
[16], or silica-supported chromium (VI) oxide in the Phillips process [17]. Both pro-
cesses are highly exothermic. A simplified free-radical mechanism is illustrated by
Fig. 16.5. Reaction conditions are set to maximize radical addition (Reactions 8 and
9) and minimize cyclization (Reactions 10 and 11).
HDPE’s are stronger than LDPE’s, but LDPE’s are cheaper and easier to make.
HDPE’s have a low degree of branching and thus stronger intermolecular forces and
tensile strength. HDPE’s can be found in many plastic containers and packaging,
such as milk jugs, detergent bottles, margarine tubs, garbage containers and water
pipes. One third of all toys are manufactured from HDPE. LDPE’s are used for both
rigid containers and plastic film applications, such as plastic bags and film wrap.
Ziegler-Natta polymerization can also produce LDPE via incorporation of an
alkyl-branched monomer, which reacts with ethylene to form a co-polymer with
short hydrocarbon branches.
The density of HDPE is greater than or equal to 0.941 g/cm3 , and the density
of LDPE falls in the range of 0.910–0.940 g/cm3 . The molecular weight of linear
polyethylene typically ranges from 200,000 to 500,000, but it can be made higher.
Using metallocenes as polymerization catalysts, manufacturers can produce ultra-
high molecular weight polyethylenes (UHMWPE) with molecular weights in the
millions, usually between 3.1 and 5.67 million [18, 19]. The densities are less than
for HDPE (for example, 0.930–0.935 g/cm3 ). They are used in a diverse range of
applications. These include can and bottle handling machine parts, moving parts on
weaving machines, bearings, gears, artificial joints, edge protection on ice rinks, and
butchers’ chopping boards.
There are many other types of polyethylenes, including linear low-density
polyethylenes (LLDPE) and cross-linked polyethylenes (PEX or XLPE). The den-
sity range of LLDPE is 0.915–0.925 g/cm3 . LLDPE’s are essentially linear polymers
with significant numbers of short branches. They commonly are made by copolymer-
ization of ethylene with short-chain alpha olefins (for example, 1-butene, 1-hexene
and 1-octene) [20].
358 16 Petrochemicals
If oxygen is added to ethylene, it reacts to form ethylene oxide. Ethylene oxide is used
to make surfactants and detergents in the manufacturing supply chain for cleaning
products. To take it even further, if water is added to ethylene oxide, it reacts in a
process called hydrolysis and forms the product ethylene glycol. Ethylene glycol is
used as an ingredient in antifreeze. It is also used to make a fiber called polyethylene
terephthalate, otherwise known as polyester, which can be found under the brand
name Dacron® . Polyester resin is also used to make beverage bottles and a whole
plethora of other useful products.
In earlier times, the production of ethylene oxide involved a chlorohydrin process,
which was less efficient than the oxidation process that is used presently. Ethylene
oxide is presently industrially produced by direct oxidation of ethylene with pure
oxygen in the presence of a silver catalyst.
Ethylene oxide is rather stable in water, but in the presence of a small amount
of acid, such as diluted sulfuric acid, it immediately forms ethylene glycol at room
temperature. The reaction also occurs in the gas phase, in the presence of a phosphoric
acid salt as a catalyst.
The reaction of ethylene oxide with water does not only end up with ethylene
glycol, it can continue to form diethylene glycol (DEG), triethylene glycol (TEG),
etc. The formation of these higher glycols is inevitable because ethylene oxide reacts
faster with ethylene glycols than with water. The most important variable is the
water-to-oxide ration, and in commercial plants the production of DEG and TEG
can be reduced by using a large excess of water.
Polyethylene glycol (PEG) is produced by the interaction of ethylene oxide with
water, ethylene glycol, or ethylene glycol oligomers, catalyzed by either acidic or
basic catalysts. The reactions with ethylene glycol and its oligomers are preferable
to produce polymers with a low polydispersity, i.e., narrow molecular weight distri-
bution. Polymer chain length depends on the ratio of reactants.
360 16 Petrochemicals
Low molecular weight PEG’s are prepared using alkali catalysts, such as sodium
hydroxide, potassium hydroxide or sodium carbonate. High molecular weight PEG’s,
or polyethylene oxides, are synthesized by suspension polymerization to hold the
growing polymer chain in solution. The catalysts are magnesium-, aluminum-, or
calcium-organoelement compounds.
PEG has been designated as nontoxic and is approved by Federal Drug Adminis-
tration (FDA) for applications in pharmaceutical formulations, foods and cosmetics.
Polyethylene glycol 3350 is in a class of medications called osmotic laxatives to treat
occasional constipation. PEG is also used in biomedical research.
The major industrial route of producing vinyl acetate involves the reaction of ethylene
and acetic acid with oxygen in the presence of a palladium catalyst.
Vinyl acetate can be polymerized to give polyvinyl acetate (PVA). PVA emulsions
in water are used as adhesives for porous materials, particularly for wood, paper, and
cloth, and as a consolidant for porous building stone, in particular sandstone [21].
With other monomers vinyl acetate can be used to prepare copolymers such as
ethylene-vinyl acetate (EVA), vinyl acetate-acrylic acid (VA/AA), polyvinyl chloride
acetate (PVCA), etc.
Vinyl chloride is made from ethylene, oxygen and hydrogen chloride (oxychlorina-
tion) over a copper(II) chloride catalyst. It can also be produced by ethylene with
chlorine over an iron(III) chloride catalyst or by thermal cracking of ethylene dichlo-
ride at 500 °C under 15–30 atm (1.5–3 MPa) pressure. Vinyl chloride is a gas with
sweet odor. It is highly toxic, flammable and carcinogenic, and therefore must be
handled with care.
Upon the action of an initiator, the double bond of vinyl chloride is opened, result-
ing in thousands of monomer units linked together to form a polymer. Polyvinyl chlo-
ride (PVC) is the third-most widely produced plastic, after PE and polypropylene
(PP). It’s used in construction for pipe and profile applications [22]. By adding plas-
ticizers, mostly phthalates, it becomes flexible for clothing and upholstery, electric
cable insulation, inflatable products and rubber replacement.
16.3 Propane, Butane, Propylene and Butylenes 361
Propane and butanes are by-products of natural gas processing and petroleum refin-
ing. The processing of natural gas involves the removal of non-methane natural gas
liquids (NGL) including ethane, propane, butanes, condensate, and natural gasoline
from the raw natural gas, in order to prevent condensation of these volatiles in natu-
ral gas pipelines, which would reduce capacity and damage the turbine compressors.
Ethane, propane, and butanes (normal butane and isobutane) are separated in a gas
processing plant through a fractionation train consisting of three distillation towers
in series: a deethanizer, a depropanizer and a debutanizer. The overhead product from
the deethanizer is ethane and the bottoms are fed to the depropanizer. The overhead
product from the depropanizer is propane and the bottoms are fed to the debutanizer.
The overhead product from the debutanizer is a mixture of normal and iso-butane,
and the bottoms product is a C5 + mixture. Isobutane can be further separated from n-
butane using a deisobutanizer. Propane and butanes are also by-products of cracking
petroleum into gasoline and cycle oils in the FCC and hydrocracking processes.
Propane can be liquefied as liquefied petroleum gas (LPG) and sold in the market
for heating, lighting and cooking. It can also be an important feedstock for making
propylene. Normal butane can be blended into gasoline to raise the Reid vapor pres-
sure of gasoline during winter in cold weather. It can be isomerized into isobutane,
which is an alkylation feed as discussed in Chap. 10.
Propylene can be made from dehydrogenation of propane, with hydrogen as a by-
product. Propylene also comes from cracking processes in the refinery. Propylene
is the second most important feedstock in the petrochemical industry. It is the raw
material for a wide variety of chemical products, shown in Fig. 16.6.
A mixture of butylenes (isobutylene, butene-1 and butene-2) is a by-product of
steam cracking of naphtha or gas oils in the manufacture of ethylene. These other
light paraffins and olefins can also be minor components of refinery fuel gas.
Propylene, butylene and isobutylene are important chemical feedstocks, with rep-
resentative products summarized in Fig. 16.6 for propylene and butylenes.
Polypropylene (PP), also known as polypropene, is the second most used thermo-
plastic polymer, after polyethylene. The structure of polypropylene has a similar
backbone as polyethylene with the methyl groups attached to every other carbon.
Different polypropylenes have different tacticities. A Ziegler-Natta catalyst is able
to restrict the linkage of monomer units to a specific regular orientation, either iso-
tactic, when all methyl groups are positioned on the same side with respect to the
backbone of the polymer chain, or syndiotactic, when the positions of the methyl
groups alternate. Isotactic polypropylene is made commercially with two types of
Ziegler-Natta catalysts, which will be described in the next section. The first group
of catalysts encompasses solid (mostly supported) catalysts or soluble metallocene
catalysts containing Ti, Zr, or Hf with an organoaluminum co-catalyst. Isotactic
macromolecules coil into a helical shape to line up next to one another to form the
crystals. Another type of metallocene catalysts produces syndiotactic polypropylene.
These macromolecules also coil into helices (of a different type) and form crystalline
materials.
16.3 Propane, Butane, Propylene and Butylenes 363
Isopropyl alcohol (IPA) is easily synthesized from the reaction of propylene with
sulfuric acid, followed by hydrolysis of isopropyl sulfate. IPA can also be produced
by reacting propylene with water, either in gas phase or liquid phase, at high pres-
sure over a solid or supported catalyst. IPA forms an azeotrope with water. Simple
distillation yields 88% IPA and 12% water. Azeotropic distillation with agents such
as diisopropyl ether or cyclohexane is required for getting pure IPA.
IPA can be oxidized to acetone using oxidizing agent, such as chromic acid or
by dehydrogenation over Raney nickel or a mixture of copper and chromium oxide.
Both IPA and acetone are important solvents in the industry. They are important
intermediates for making a variety of chemicals. IPA has been used as a mixture
with water as rubbing alcohol and other medical sanitary purposes.
Acrylic acid is produced by two stages of vapor oxidation of propylene over catalyst.
In the first stage, propylene is oxidized to acrolein over a highly active and very
selective catalyst (V or Mo oxides). In the second stage, acrolein is further oxidized
over a Co-Mo catalyst at 200–300 °C.
16.3 Propane, Butane, Propylene and Butylenes 365
Acrylic acid is miscible with water, alcohols, ethers and chloroform. Acrylic acid
is a precursor of a wide variety of chemicals in the polymer and textile industries.
Examples are manufacture of plastics, latex applications, in floor polish, in poly-
mer solution for coating applications, emulsion polymers, paint formulation, leather
finishing and paper coating. Acrylic acid is used to make acrylic esters and resins,
chemicals added to protective coatings and adhesives, and the production of poly-
acrylic acid polymers.
Other than oxidation of propylene, acrolein can also be obtained by thermal
decomposition of glycerol at 280 °C. Acrolein is mainly used as a contact herbi-
cide to control submersed and floating weeds, as well as algae, in irrigation canals.
In the oil and gas industry, it is used as a biocide in drilling fluids, as well as a
scavenger for hydrogen sulfide and mercaptans.
Propylene can be oligomerized to trimers, tetramers and other olefins in the presence
of acids. Nonylphenols are produced by acid-catalyzed alkylation of phenol with a
mixture of propylene trimers. They are a family of closely related organic compounds
called alkylphenols, and used in manufacturing antioxidants, lubricating oil additives,
detergents, emulsifiers, cosmetics, and insecticides. These compounds are important
non-ionic surfactants alkylphenol ethoxylates and nonylphenol ethoxylates, which
are used in detergents, paints, pesticides, personal care products, and plastics.
The reactants pass through a fluidized bed reactor containing bismuth phospho-
molybdate supported on silica at 410–510 °C and 50–200 kPa. The reaction product is
quenched by aqueous sulfuric acid, yielding a mixture containing acrylonitrile, ace-
tonitrile, hydrocyanic acid and ammonium sulfate. After removal of water-soluble
components, acrylonitrile and acetonitrile are separated by distillation.
16.3 Propane, Butane, Propylene and Butylenes 367
Acylonitrile (ACN) is highly flammable and toxic. It’s also suspected to cause
lung cancer through exhaust emission and cigarette smoking.
ACN is used mainly as a monomer or comonomer in the production of synthetic
fibers, plastics and elastomers. It is the monomer source of polyacrylonitrile (PAN), a
homopolymer, and precursor of high quality carbon fiber; styrene-acylonitrile (SAN);
acylonitrile butadiene styrene (ABS); acrylonitrile styrene acrylate (ASA) and other
rubbers, such as acrylonitrile butadiene rubber (NBR).
Nitrile rubbers are used in the automotive and aeronautical industries to fabricate
hoses for fuel and oil handling, and to make seals and grommets. Other uses include
gloves, footware, adhesives, sealants, sponges, floor mats, etc.
Acrylonitrile is also a precursor in the industrial manufacture of acrylamide and
acrylic acid.
Higher carbon number alcohols can be manufactured from lower carbon number
olefins with carbon monoxide and hydrogen through hydroformulation via the oxo
process, which is discussed in Sects. 16.4 and 16.5.
16.3.11 Polybutyenes
Butyl rubber [27, 28] was invented in 1937 by Standard Oil of New Jersey (now
ExxonMobil). During World War II, this synthetic rubber enabled Canada and the
United States to manufacture tires after supplies of natural rubber from Southeast
Asia were cut off. Butyl rubber is a copolymer of isobutylene with isoprene, pro-
duced by polymerization of about 98% of isobutylene with about 2% of isoprene for
crosslinking by vulcanization, with the following representative structure. It is imper-
meable to air and used in many applications requiring an airtight rubber.
PIB and butyl rubber are also used in adhesives, agricultural chemicals, fiber optic
compounds, ball bladders, caulks and sealants, cling film, electrical fluids, lubricants
(2-cycle engine oil), paper and pulp, personal care products, pigment concentrates,
for rubber and polymer modification, for protecting and sealing certain equipment for
16.3 Propane, Butane, Propylene and Butylenes 369
use in areas where chemical weapons are present, as a gasoline/diesel fuel additive,
and even in chewing gum.
As a fuel additive, PIB has detergent properties. When added to diesel fuel, it
resists fouling of fuel injectors, leading to reduced hydrocarbon and particulate emis-
sions. It is blended with other detergents and additives to make a “detergent package”
for gasoline and diesel fuel to resist buildup of deposits. Such deposits can catalyze
pre-ignition, so preventing their formation reduces engine knock.
Butyl rubber and halogenated rubber are used for the inner liner (“inner tube”)
that holds air in tires. Bromobutyl has air impermeability and resists UV radiation
properties which make it ideal for tubeless tires.
16.3.14 Butadiene
CO/H2
HCo(CO)4
The most common process for making butanol starts with propylene, which under-
goes hydroformylation to form normal- and iso-butyraldehydes, which are then
reduced with hydrogen to 1-butanol and/or 2-butanol. Secondary and tertiary butanols
can be made from butane-1 and isobutylene, respectively, with water in the presence
of sulfuric acid. Butanol can also be produced by fermentation of biomass by bacte-
ria (as biobutanol).
16.5 Butanols and Methyl Ethyl Ketone (MEK) 371
Benzene can’t be used in gasoline due to health and environmental concerns, but
other reformer products—toluene, xylenes, and EB—are high-octane blend stocks.
Planners often must decide, based on prices and relative demand, whether to blend
these products into gasoline or send them to a chemicals plant.
The reaction of benzene with propylene yields cumene, which produces a mixture
of acetone and phenol via cumene peroxide. The condensation of phenol and acetone
in a 2:1 molar ratio produces bisphenol A (BPA). BPA is a starting material for the
synthesis of plastics, primarily polycarbonates for drinking water bottles and epoxy
resins.
Nonylphenol is produced by the acid catalyzed alkylation of phenol with a mixture
of nonenes. Nonyl phenol is used in manufacturing antioxidants, lubricating oil
additives, laundry and dish detergents, emulsifiers and solubilizers.
Partial hydrogenation of phenol or hydrogenation of benzene gives cyclohex-
ane which yields cyclohexanol and cyclohexanone upon oxidation. The mixture of
cyclohexanol and cyclohexanone is oxidized with nitric acid to give adipic acid. The
reaction of the acid with hexamethylenediamine produces Nylon 66, a widely used
fiber and plastic.
16.6.1.1 Nylons
Nylon is a thermoplastic, silky material and was the first commercially successful
synthetic thermoplastic polymer produced by DuPont in the 1930s. Commercially,
Nylon 66 is made by reacting adipic acid with hexamethylenediamine, shown on
top of Fig. 16.9. Another polymer of Nylon 6 is made from only one monomer of
caprolactam through ring opening polymerization, shown in the bottom of Fig. 15.9.
Both adipic acid and caprolactam can be derived from benzene through cyclohex-
ane and cyclohexanone/cyclohexanol. The most common manufacturing process of
adipic acid is the nitric acid oxidation of a cyclohexanol and cyclohexanone mixture,
called KA (for ketone-alcohol) oil.
Cyclohexane is produced from benzene by liquid phase hydrogentation at a pres-
sure of 1.34 MPa and a temperature of 210 °C using a Raney nickel catalyst. The
mixture of cyclohexanol and cyclohexanone is produced by continuous oxidation of
cyclohexane by air using a trace of cobalt naphthenate as catalyst.
Caprolactam is produced from cyclohexanone by being converted into oxime
with hydroxylamine first, which is then treated with oxime with acid to introduce
Beckmann rearrangement for yielding ε-caprolactam.
16.6 BTX Derivatives 373
TNT is one of the most commonly used explosives for military, industrial, and
mining applications. TNT has been used in conjunction with hydraulic fracturing.
Toluene diisocyanate (TDI), CH3 C6 H3 (NCO)2 , has six possible isomers. Two of
them are commercially important: 2,4-TDI and 2,6-TDI. 2,4-TDI are prepared in
three steps from toluene via dinitrotoluene and 2,4-diaminotoluene (TDA). Finally,
the TDA is treated with phosgene to form TDI, with HCl as a co-product that is a
major source of industrial hydrochloric acid.
Distillation of the raw TDI mixture produces an 80:20 mixture of 2,4-TDI and
2,6-TDI, known as TDI (80/20). Differentiation or separation of the TDI (80/20) can
be used to produce pure 2,4-TDI and a 65:35 mixture of 2,4-TDI and 2,6-TDI, known
as TDI (65/35) [32]. All TDIs are reactive chemicals and potentially hazardous to
humans.
TDIs are important building blocks to produce polyurethanes by reacting with
hydroxy groups, such as a wide range of polyols, with formation of carbamate links.
Polyurethanes are used to produce countless products, particularly in the transporta-
tion and construction industries, such as rigid and flexible thermoset foams, coatings,
adhesives, sealants, elastomers and lighter automobile parts for energy savings.
There are three xylene isomers—ortho, meta, and para. The most valuable is puri-
fied p-xylene, which undergoes oxidation to form terephthalic acid (TPA), shown
below. TPA is a fundamental building block for polyethylene terephthalate (PET),
commonly known as polyester.
Several refinery processes are devoted to converting other small aromatics into
p-xylene. The UOP Parex process recovers high-purity p-xylene from mixtures of C8
aromatics [35]. In the UOP Isomar process [36], p-xylene-free mixtures of o-xylene
and m-xylene are allowed to equilibrate, forming a new mixture which contains all
three isomers along with ethylbenzene. At one set of conditions, an equilibrium mix-
ture contains 60% m-xylene, 14% p-xylene, 9% o-xylene, and 17% EB. Therefore, it
is important that the Isomar process also converts ethylbenzene by disproportionation
into benzene and xylenes. EB itself can also be converted to styrene (vinyl benzene)
by dehydrogentation. Styrene is then polymerized to form polystyrenes, one of the
most widely used plastics.
Phthalic anhydride is a principal commercial form of phthalic acid. It is produced
from oxidation of o-xylene over a vanadium pentoxide (V2 O5 ) catalyst between
320–400 °C. Phthalic anhydride is an important ingredient to make plasticizers with
oxo alcohols.
376 16 Petrochemicals
As mentioned, synthesis gas (syngas) is a mixture of CO and hydrogen and very often
some carbon dioxide. CO is produced commercially by partial oxidation of natural
gas, coal or biomass. Both CO and H2 can be produced by steam/methane or naphtha
reforming. Additional H2 can be obtained from the water-gas shift reaction, with CO2
as a byproduct. Many of the reactions derived from these gases have been discussed
in Sect. 16.1. The main use of syngas is to produce alkanes, olefins and alcohols
through the Fischer-Tropsch process. In South Africa during apartheid, when due to
an embargo the country could not import enough petroleum to meet its needs, syngas
from coal was the main source of commercial hydrocarbons. A similar situation
developed in Germany during World War II. The country supplemented its limited
supplied of petroleum by converting coal into syngas, which in turn was converted
through Fischer-Tropsch synthesis into fuels and lubricants for war machines (planes,
tanks, ships, etc.). Syngas can also be produced from biomass for the production of
renewable biofuels and biochemicals.
Liquid transportation hydrocarbon fuels and various other chemical products can
be produced from syngas via Fischer-Tropsch synthesis (FTS), named after the two
German inventors, Franz Fischer and Hans Tropsch, in the 1920s. The Fischer–Trop-
sch (F-T) process involves a series of chemical reactions that produce a variety of
hydrocarbons:
where n is typically 10–20. The alkanes are linear and fall in the diesel range. Higher
hydrocarbons are produced at low temperatures (150–300 °C) with cobalt catalysts.
In addition to alkane formation, competing reactions give small amounts of alkenes,
as well as alcohols and other oxygenated hydrocarbons. The technology is often
referred to as a gas-to-liquid (GTL) process for natural gas.
16.7 Synthesis Gas (Syngas) 377
Wn /n → (1 − α)2 αn−1
Hydrodewaxing
Hydroisomerization
Syngas Hydrocracking
CO + H2
Diesel
High Temperature FTS
Fe Catalyst Olefins (C3 - C11)
Fluidized Bed Reactor
Oligomerization
Isomerization
Reforming
Hydrogenation
Gasoline
Fig. 16.11 Various fuel and lubricant products from high-temperature and low-temperature
Fischer-Tropsch synthesis (FTS) routes
The syngas to gasoline plus (STG+) process consists of four fixed bed reactors in
series in which syngas is converted to synthetic fuels. For producing high-octane
gasoline, syngas is fed to first reactor to convert most of the syngas (CO and H2 )
to methanol by passing through the catalyst bed. The methanol-rich gas enters the
second reactor to be converted to dimethyl ether (DME) by catalytic dehydration.
The gas from the second reactor is next fed to the third reactor containing a catalyst to
convert DME to hydrocarbons, including paraffins, naphthenes, aromatics and small
amounts of olefins. Then, the fourth reactor provides transalkylation and hydrogena-
tion to reduce durene/isodurene and trimethylbenzene leading to desirable viscomet-
ric properties. Finally, the mixture enters a separator to separate non-condensed gas
from liquid product. The gas is recycled back to the first reactor.
16.7 Synthesis Gas (Syngas) 379
Methanol is produced when the CO and CO2 produced in steam methane reforming
and water shift reactions are reacted with hydrogen in a ratio of CO2 :CO:H2 = 5:5:90
at 50–100 bars and 225–275 °C over Cu/ZnO/Al2 O3 catalysts. The predominant
reactions are:
Syngas can also be produced from integrated gasification combined cycle (IGCC)
where a high-pressure gasifier with sub-stoichiometric oxygen turns heavy petroleum
residue, coke, coal and biomass into pressurized gas. Impurities such as sulfur diox-
ide, particulates, mercury, and in some cases carbon dioxide are removed prior to
power generation. Syngas cleanup includes the removal of bulk particulates by filters,
fine particulates by scrubbing and mercury by solid adsorbents. Carbon monoxide
emission can be reduced by adding a water-gas shift reactor. Carbon dioxide for the
shift reaction can be separated, compressed and stored through sequestration. After
the shift reactor, the syngas is fairly pure hydrogen.
The syngas and hydrogen are used for power generation to drive generators rather
than making chemicals. The IGCC plant is called integrated because (1) the syngas
produced in the gasifier is used as fuel for the gas turbine and (2) the steam generated
by syngas cooler in the gasifier is used to drive a steam turbine in the combined
cycle for electricity generation. A third function is sometimes added: co-generation
of hydrogen for refineries and other industrial facilities. This process could increase
values for undesirable cokes that cannot be sold in the market.
380 16 Petrochemicals
Refineries supply several feedstocks for petrochemical plants (Table 16.1). Primary
petrochemicals can be divided into three groups: olefins, aromatics, and synthesis
gas.
As mentioned, olefins are major sources of polyolefins, such as polyethylene,
polypropylene, and ethylene/propylene co-polymers. Ethylene and propylene are
sources of other polymers, too. Ethylene reacts with benzene to make ethyl benzene,
which leads to styrene and polystyrene. Propylene is oxidized to acrylic acid, used
to make polyacrylates. Oxidation of propylene in the presence of ammonia produces
acrylonitrile, which leads to polyacrylonitrile and several derivatives. Butyl rubbers
made from C4 olefins are basic ingredients for tires. Bromobutyl (and Chlorobutyl)
is the key ingredient for tubeless tires due to its excellent air impermeability, UV
resistant and antioxidation properties. High polymers can be further processed into
plastics and engineering plastics.
Aromatics from catalytic reformers are mainly benzene, toluene and xylenes.
They are solvents and feedstocks for a variety of chemicals. p-Xylene is a major feed
for polyester production.
Synthesis gas is used for making methanol and ammonia as well as synthetic fuels
and lubricants.
Figure 16.12 illustrates that it’s a two-way street: petrochemical plants also feed
refineries. The olefin plant is essentially a steam cracking plant to produce olefins.
Hydrogen is, for the most part, a by-product of olefins plants, but it has high value
in refineries. Pyrolysis gasoline (py gas) from steam cracking can be stabilized by
hydrotreating and blended into the gasoline pool. Similarly, excess butanes from
olefins plants can be converted into high-value gasoline blendstocks or into isobutane
as a feedstock for alkylation in refineries.
Feedstocks make up anywhere from 60 to 70% of the cost to manufacture petro-
chemicals; therefore, the feedstock flexibility of domestic petrochemical processes
has enabled US producers to remain competitive. As both oil and gas became more
expensive in the US during the early 2000s, however, coupled with increased energy
Ethane,
Gas processing
propane
Ethylene
Naphtha
hydro- PC naphtha Propylene (chemical grade)
treater
Upgraded
bottoms
Pygas Hydrogen
Benzene
Refinery
hydrocracker To refinery
C4 compounds
Butadiene
C4 processing Butane
Butenes
Refinery FCC To refinery
Xylenes
alkylation
Refinery Benzene
catalytic reformer
Cycloexane
Benzene hydrotreater
costs, the competitive advantage that came with feedstock flexibility all but disap-
peared. Petrochemical companies began to seek joint ventures in other parts of the
world where energy and feedstock costs were cheaper. All that changed around 2008,
when production of oil and gas from shale “fracking” led to a dramatic decrease in
the cost of industrial energy and feedstocks in the US.
Shale development is providing manufacturers with an unprecedented opportunity
in low cost energy; however, that is only part of the picture. Along with the natural
gas and oil coming from shale plays like the Marcellus, Bakken, Utica and Eagle
Ford, these areas contain an abundance of NGLs. Ethane is a major component of
NGLs, especially in the Marcellus, Utica and Eagle Ford plays. The economic law of
supply and demand states that an increase in supply, other things being equal, results
in a decrease in price. This new-found bounty of NGLs has significantly reduced the
price of ethane in the US. The low cost of energy and ethane has prompted American
petrochemical manufacturers to expand ethane cracking capacity. Over the past three
years, petrochemical companies have announced plans to invest over $80 billion in
new manufacturing infrastructure.
382 16 Petrochemicals
References
The petroleum industry is often divided into three business sectors: Upstream, Mid-
stream and Downstream [1, 2]. Upstream and downstream have been well defined
and adopted in petroleum companies and communities.
Upstream involves the discovery (exploration) and recovery (production) of
petroleum resources, including natural gas, condensate, and crude oil, from under-
ground or underwater fields. The recently developed in situ steam-assisted gravity
drainage (SAGD) and hydraulic fracturing (fracking) to recover trapped gas and oils
from shales and other tight formations are upstream operations. The open mining of
oil sand bitumen deposits belongs to upstream.
On the other hand, downstream operations include the refining of crude oils, pro-
duction of petrochemicals from petroleum, Fischer-Tropsch production of synthesis
gas (syn gas), and distribution of petroleum-based products. Refining produces fuels,
lubricant oils, and petrochemicals. Petrochemicals include light olefins and BTX
(benzene, toluene and xylenes), steam-cracker feedstocks such and light naphtha
and hydrocracker unconverted oil. In turn, petrochemicals are converted into thou-
sands of chemicals, including solvents, polymers, fibers, coatings, plastics, synthetic
rubbers, and pharmaceuticals.
Generally speaking, “midstream” connects upstream with downstream. Mid-
stream operations and processes generally include transportation by pipeline, tanker,
barge, rail and trucks; storage; and trading of crude oil and refined products [2].
Trading of refined products is often considered to be a downstream business. Trad-
ing natural gas, crudes and other refinery feedstocks can be included in upstream.
The midstream sector is huge. Estimates of total pipeline mileage differ widely.
According to the United States Department of Transportation, in 2014 the United
States had the largest oil and gas pipeline network in the world, accounting for
~40% of the world total [3]. According to Pipeline 101, in 2018 the U.S. has more
than 2.4 million miles of gas pipeline pipelines and more than 190,000 miles of
liquid petroleum pipelines, of which 72,000 miles are for crude oils [4]. The cost of
installing a pipeline ranges from $300,000 to $1.5 million per mile, so the cost of
replacing this infrastructure would be enormous. The United States Central Intelli-
gence Agency reports that there are 4295 ocean-going oil tankers in the world, each
with a deadweight tonnage (DWT) greater than 1000 long tonnes [5].
“Midstream,” generally starts after wellhead treatments or at the point where oil,
gas, or other produced hydrocarbons enter a transportation system—a pipeline or a
vehicle, such as a barge, tanker or rail car.
There are some gray areas. The liquefaction of natural gas to LNG, extrac-
tion/dilution of bitumen, and production of synthetic crude oil (syn crude) for trans-
portation seem to fit best into midstream rather than upstream. The processing of
natural gas includes recovering sulfur, carbon dioxide, helium, and hydrocarbon liq-
uids before the gas goes to a product pipeline. CO2 will form dry ice upon refrigeration
to clog tubes in pipelines and tankers. Hence, it has to be removed prior to natural
gas liquefaction for transportation. The acid gases (CO2 and H2 S) are removed with
alkanol amines, such as diethanol amine (DEA); usually, the H2 S is carried by alka-
nol amine to a Claus plant, where it is converted into sulfur. The physical separation
of gas, oil, water and particulates in a sedimentation vessel at or near well head,
however, is an upstream process.
The dilution of bitumen by froth treatment, and the upgrading of the oil sand
bitumen into synthetic crude oil via coking prior to pipeline transportation, is called
midstream operations by some companies and downstream operations by others. Salt
dome storage of natural gas or crude oil, including in strategic reserves, can also be
considered as a midstream operation.
Mid-stream ends at the point where a hydrocarbon stream enters refinery storage
tanks (i.e., the tank farm). Hence, storage tank management at a transshipment point
is part of midstream, but refinery tank farms are a part of downstream operations,
including crude oil blending.
17.2 Transportation
Large quantities of crude oils are transported by pipelines, either on or beneath land
or underwater, or in tankers. Seagoing tanker transportation has been heavily used
where pipeline transportation becomes difficult, uneconomical or impossible, such
as for intercontinental transportation. The larger the tanker, the lower is the unit
cost of transportation. Viscous oil sand bitumen produced in Canada is diluted by
condensates, naphtha or middle distillates, or converted to synthetic crude oil on-site
for pipeline transportation.
For transportation of crude oil through pipeline, the excess quantities of water
have to be removed to meet specification. The maximum tolerable water content for
pipeline transportation is from 0.5 to 2.0%.
17.2 Transportation 387
Fig. 17.1 Transportation of natural gas, crude oils and their products [6]
Figure 17.1 summarizes transportation means for crude oil and natural gas as
well as their products. Not shown are the storage facilities (tank farms) needed at
transportation ports and refining centers for the large quantities of petroleum. Crude
oils may also be stored in such geological features as salt domes, used by the U.S.
Strategic Petroleum Reserve.
Pipelines are the most convenient and economical means of transporting crude oil
from the producing field to the refinery. For long distance or over a vast of water
body, such as an ocean, where construction of pipelines becomes impractical, tankers
are used. However, pipelines are still needed for transporting the oils to marine
terminals for loading onto tankers and unloading at destination ports for shipping to
the refineries.
In recent years, heavy oil, extraheavy oil and bitumen resources are more than dou-
ble the conventional oil reserves worldwide. Very large reserves are mainly located in
Venezuela and Canada. Because of their extremely low mobility due to high viscosity
(>1000 cP), prior reduction in viscosity is necessary for transportation via pipeline.
In addition to the flow properties, asphaltene deposition, heavy metals, sulfur and
388 17 Midstream Transportation, Storage, and Processing
Fig. 17.2 Technologies of transporting heavy oil and bitumen through pipeline [7]
brine or salt content present other challenges. Hence, upgrading heavy crude oil and
bitumen to meet conventional light crude oil properties become an option.
Common technologies of transporting heavy oil and bitumen through pipeline
include preheating, blending or dilution, emulsification, partial upgrading and core-
annular flow, summarized in Fig. 17.2. Each of these techniques is aimed at reducing
viscosity as well as energy required for pumping, to enhance flowability of the oil
via pipelines.
For viscosity reduction, dilution/blending is a commonly used technique. The
most widely used diluents include condensate, naphtha, kerosene and light crude oil.
The resulting blend of heavy crude oil and diluents, i.e., diluted bitumen or “dilbit”,
has lower viscosity and therefore easier to pump. However, instability of the blend
due to asphaltene precipitation should be avoided by choosing the proper diluent.
Preheating followed by subsequent heating of the pipeline enhances the flow
properties of heavy crude oil and bitumen. But heating pipelines is expensive due to
high heat loss, a greater number of pumping/heating stations, expansion of pipelines,
and greater corrosion inside the pipe due to the higher temperature. Consequently,
heated pipes are used only for short distance transportation within a processing
facility.
The emulsion method consists of dispersing heavy crude oil in water in the form
of droplets stabilized by surfactants. The surfactants stabilize the emulsion to prevent
drop growth and phase separation.
The use of polymers (polyacrylates, polymethacrylates, etc.) as pour point depres-
sants inhibits precipitation of high density and high viscosity asphaltenes.
Drag-reducing additives, such as silicones, polymers, fibers and surfactants sup-
press turbulent flow and reduce friction near the pipeline wall. In core-annular flow,
the heavy crude oil flows through the pipeline with a film layer of water or solvent
near the pipe wall, which acts as a lubricant, thus, reducing the pump pressure drop
17.3 Transportation of Heavy Oil and Bitumen 389
over a long distance as comparable to the one for water alone. Due to their relatively
low cost and ease of handling, silicones are the most widely used friction reducers,
but they are problematic because they poison refinery catalysts.
In situ upgrading is achievable by thermal recovery of the oil and bitumen, includ-
ing SAGD, CSS, THAI, etc. These upstream processes involve thermal cracking
underground, which yield lighter hydrocarbons with lower viscosity.
Oil sand or tar sand is a mixture of bitumen, sand, water and clay. It does not act like
conventional crude oil. It must be mined in an open pit or in situ underground. Then
its viscosity must be reduced prior to transportation to refineries for upgrading into
finished products. For open mining, froth treatment is needed to separate bitumen
from the aqueous and solid contaminants (sand, water and clay) [8]. This extraction
process is not required for in situ processing, such as SAGD (Chap. 7, Sect. 7.9.2.3),
because the separation happens in the ground.
In surface-mined oil sands operations, bitumen is extracted from oil sand ore using
a warm-water extraction process that produces bitumen froth typically containing
60 wt.% bitumen, 30 wt.% water, and 10 wt.% mineral solids. The froth is first
diluted with a naphthenic or paraffinic hydrocarbon solvent to reduce the viscosity
and density of the oil phase, thereby accelerating the settling of the dispersed phase
impurities by gravity or centrifugation. Figure 17.3 shows the froth formation after
mixing bitumen with solvent for 2, 4, 8, 12 and 16 min. The mineral and inorganic
impurities settled down to the bottom. The diluted product is called “diluted bitumen”
Synthetic crude oil is a blend of naphtha, distillate, and gas oil range materials,
with little residuum 1050 °F + (565 °C+) material. It is a high-quality intermediate
product from a bitumen/extra heavy oil upgrader facility used in connection with oil
sand production. Synthetic crude oil also is produced from oil shale, which comes
from kerogen-bearing rock. Oil in shale is recovered by pyrolysis (retorting) [14].
The properties of the synthetic crude depend on the processes used in the upgrading.
Typically, it is low in sulfur and has an API gravity of around 30, with properties
close to conventional crude oil.
In the syncrude plant, the bitumen is heated and sent to drums where excess
carbon, in the form of petroleum coke, is removed. The superheated hydrocarbon
vapors from the coke drum are sent to fractionators where vapors condensed into
naphtha, kerosene and gas oil. These products are blended into synthetic crude oil
which is shopped by pipeline to refineries. Diesel can be sold as a final product. Shale
oil products often contains significant amounts of arsenic, which is another severe
catalyst poison in refineries [15].
The largest producers of synthetic crude in the world are Syncrude Canada, Suncor
Energy Inc., and Canadian Natural Resources Limited, with a cumulative production
of approximately 600,000 barrels per day (95,000 m3 /d).
Much natural gas comes from production sites far from the market place. Natural
gas is mostly methane. After removal of water, H2 S, CO2 , mercury, and higher
hydrocarbons (condensate), natural gas is transported by pipeline or by tankers. For
tanker transport„ natural gas is converted into liquefied natural gas (LNG), a process
called liquefaction.
Figure 17.5 shows the sequence of transportation of LNG. Transportation from
the field to the liquefaction facility is accomplished with gas pipelines. Treatment in
the facility involves the removal of acid gases, H2 S and CO2 , and mercury. The con-
densate and water are also removed; condensate is a valuable byproduct. Precooling
392 17 Midstream Transportation, Storage, and Processing
removes heavy hydrocarbons, which can be combined with condensate for storage.
The remaining gas is refrigerated. The refrigerant gas is compressed, cooled, con-
densed and dropped in pressure through a valve to reduce temperature by the Joule-
Thomson effect. The temperature of the feed gas is eventually reduced to −161 °C.
At this temperature, all hydrocarbon gas including methane liquefies. Constituents
of natural gas (propane, ethane and methane) are typically used as refrigerant gases.
Since carbon dioxide normally co-exists with natural gas in gas wells, the removal
of carbon dioxide by diethanolamine (DEA) or equivalent is necessary to avoid dry
ice formation clogging the pipes, tubing, and other transferring components, such as
valves. Several options for amine treating were discussed in Chap. 14, Sect. 14.1.2.
In sour gas fields, such as those in Western Canada, both CO2 and H2 S are removed
in the same acid gas scrubbing facility.
Mercury is present in natural gas predominantly as elemental mercury. It is highly
corrosive to LNG aluminum alloy cryogenic exchangers, causing equipment failure.
Adsorption is a primary method for its removal. Sulfur impregnated carbon captures
mercury as mercuric sulfide and fixes the compound due to its high energy adsorption
into the pores of the adsorbent. Water can be removed simultaneously. The mercury
level of the treated gas can be less than 0.01 µg per normal cubic meter.
The final product is LNG, ready for storage and shipping. A flow chart for the
LNG process is shown in Fig. 17.6 [16]. Hydrogen and helium remain as gases to be
recovered. As mentioned earlier, natural gas plants are the major commercial sources
of helium.
References
1. Global Energy: Challenges and Solutions in an Upstream and Downstream Oil and
Gao Operation: http://globalenergy.pr.co/65678-challenges-and-solutions-in-an-upstream-
and-downstream-oil-and-gas-operation. Retrieved 6 Aug 2016
2. Wikepedia: Midstream. https://en.wikipedia.org/wiki/Midstream. Retrieved 6 Aug 2016
3. Wikipedia: https://www.bts.gov/content/us-oil-and-gas-pipeline-mileage. Retrieved 8 Oct
2018
4. http://www.pipeline101.org/where-are-pipelines-located. Retrieved 8 Oct 2018
5. Retrieved November 8, 2016
6. United States Central Intelligence Agency: The World Factbook, https://www.cia.gov/library/
publications/the-world-factbook/fields/2108.html
7. Speight JG (2007) The chemistry and technology of petroleum, 4th edn. CRC Press, Boca
Raton
8. Hart A (2014) A review of technologies for transporting heavy crude oil and bitumen via
pipeline. J Petrol Explor Prod Technol 4:327–336
9. https://thenarwhal.ca/environment-canada-study-reveals-oilsands-tailings-ponds-emit-toxins-
atmosphere-much-higher-levels-reported. Access 30 Nov 2018
10. https://www.cbc.ca/news/business/fort-mcmurray-greenhouse-to-turn-garbage-into-veggies-
1.2984722. Access 30 Nov 2018
11. https://www.oilsandsmagazine.com/technical/mining/froth-treatment. Retrieved 7 Oct 2018
12. Froth Treatment: https://www.nrcan.gc.ca/energy/oil-sands/5873. Natural Resources Canada.
Retrieved 15 Sept 2015
394 17 Midstream Transportation, Storage, and Processing
13. Rao F, Liu Q (2013) Froth treatment in Athabasca Oil Sands Bitumen Recovery Process: a
review. Energy Fuels 27(12):7199–7207
14. Frost CM, Paulson RE, Jensen HB (2018) Production of synthetic crude shale oil produced by
in-situ combustion retorting. https://web.anl.gov/PCS/acsfuel/preprint%20archive/Files/19_2_
LOS%20ANGELES_04-74__0156.pdf. Retrieved 7 Oct 2018
15. https://web.anl.gov/PCS/acsfuel/preprint%20archive/Files/23_4_MIAMI%20BEACH_09-
78_0018.pdf. Access 30 Nov 2018
16. https://petrowiki.org/File:Vol6_Page_369_Image_0001.png. Retrieved 7 Oct 2018
Chapter 18
Safety and Environment
18.1 Introduction
In addition to desired products, refineries and chemical plants also produce signifi-
cant amounts of toxic, harmful and environmentally damaging organic and inorganic
materials. Hence, environmental considerations are addressed during design, pro-
cessing, product handling, and disposal. Various treatment, capture and disposal units
are required. Fugitive emissions from evaporation, sedimentation, seeping, drifting,
spreading and natural dispersion are constantly monitored. Air and water qualities
after disposal have to meet specifications and standards. Oil spills into water and
onto soil threaten environment. Inadvertent discharges can cause tremendous loss
in life and environmental damages on a large scale, such as that which happened
after Texas City refinery explosion, the ExxonValdez oil spill and the recent deadly
blowout of the Deepwater Horizon in the Gulf of Mexico.
Air Quality. In the 1970s and 1980s, environmental laws compelled oil and
gas companies to reduce emissions of SOx , NOx , CO2 and hydrocarbons. In the
atmosphere, SOx reacts with water vapor to make sulfurous and sulfuric acids and
sulfate particulates, which return to earth as wet or dry acid deposition; wet deposition
is often called acid rain. Volatile hydrocarbons react with NOx to make ozone. CO2
is a major “green-house” gas. To reduce these pollutants, the industry tightened its
operations by:
• Reducing fugitive hydrocarbon emissions from valves and fittings
• Removing sulfur from refinery streams and finished products
• Adding tail-gas units to sulfur recovery plants
• Reducing the production of NOx in fired heaters
• Scrubbing SOx and NOx from flue-gases
• Reducing the production of CO2 by increasing energy efficiency.
In several oil fields, companies keep CO2 out of the atmosphere by pumping it
into oil wells to maintain reservoir pressure, thereby enhancing production.
Waste Water Treatment. Waste water treatment is used to purify process water,
runoff, and sewage. As much as possible, purified waste-water streams are re-used
in the refinery. Wastewater streams may contain suspended solids, dissolved salts,
phenols, ammonia, sulfides, and other compounds. The streams come from just about
every process unit, especially those that use steam, wash water, condensate, stripping
water, caustic, or neutralization acids.
Primary treatment uses a settling pond to allow most hydrocarbons and suspended
solids to separate from the wastewater. The solids drift to the bottom of the pond,
hydrocarbons are skimmed off the top, and oily sludge is removed. Difficult oil-in-
water emulsions are heated to expedite separation. Acidic wastewater is neutralized
with ammonia, lime, or sodium carbonate. Alkaline wastewater is treated with sul-
furic acid, hydrochloric acid, carbon dioxide-rich flue gas, or sulfur.
Some suspended solids remain in the water after primary treatment. These are
removed by filtration, sedimentation or air flotation. Flocculation agents may be
added to consolidate the solids, making them easier to remove by sedimentation or
filtration. Activated sludge is used to digest water-soluble organic compounds, either
in aerated or anaerobic lagoons. Steam-stripping is used to remove sulfides and/or
ammonia, and solvent extraction is used to remove phenols.
Tertiary treatment processes remove specific pollutants, including traces of ben-
zene and other partially soluble hydrocarbons. Tertiary water treatment can include
ion exchange, chlorination, ozonation, reverse osmosis, or adsorption onto activated
carbon. Compressed oxygen may be used to enhance oxidation. Spraying the water
into the air or bubbling air through the water removes remaining traces of volatile
chemicals such as phenol and ammonia.
Solid Waste Handling. During drilling and production operations, oil-bearing
cuttings from wells present environmental challenges. Onshore, the cuttings can be
stored onsite in clay-lined pits. Offshore, the cuttings are returned to the sea. Other
wastes are transported to waste treatment facilities. In the oil sands fields in Alberta,
Canada, bitumen is washed way from the inorganic substrate with hot water. The
water goes into huge tailings ponds. The ponds emit considerable amounts of CO2 ,
and they contain significant amounts of trace contaminants.
Refinery solid wastes may include spent catalysts and catalyst fines, acid
sludge from alkylation units, and miscellaneous oil-contaminated solids. All oil-
contaminated solids are treated as hazardous and sent to sanitary landfills.
Recently, super-critical extraction with carbon dioxide has been used with great
success to remove oil from contaminated dirt, including cuttings from oil-well
drilling.
Improving Energy Efficiency and Safety with Automation. The ever-
increasing development and use of better instruments and analyzers, engineering
models, and real-time online optimization continue to improve the efficiency and
safety of petroleum processing plants. Subsequent sections provide examples. Imple-
menting advanced process control increased production by 2.5% at the Chemopetrol
ethylene plant in Litvinov, Czech Republic. Other examples are provided on the web
sites of companies such as Applied Manufacturing Technologies, Aspen Technology,
18.1 Introduction 397
130 Standard Oil 1870 Anglo-Persian Oil 1909 Crises in Iran 1978-80 2008 Financial Crisis
120
Russian exports 1873- Zumaque 1914 Brent 1976
110
100 Telaga Said well 1885 Bahrain 1932 Prudhoe Bay 1967
US$ per barrel
Fig. 18.1 Impact of world events on crude oil prices. After the 2008 financial crisis, prices dropped
considerably due to lower consumption. Similarly, in 1973, conservation initiatives stalled the
decades long upward trend in consumption
shutdowns. The slogan for ExxonMobil’s program is Nobody Gets Hurt [1]. Shell
preaches Goal Zero [2]. KBR promotes Zero Harm 24/7 [3]. Chevron’s ten Tenants
of Operation [4] are prefaced by two statements: “Do it safely or not at all. There is
always time to do it right.” These particular companies have departments dedicated
to HSE (health, safety, environment) led by executives who report directly to the
CEO or COO, not just to someone in human resources. The 0-0-0 initiatives are
not just for show. They reflect a sincere belief that our business doesn’t have to be
deadly. HSE rules are strictly enforced. Employees who violate them are dismissed.
This relatively recent shift in attitude requires investment, but the return on such
investments is enormous in many measureable ways.
Vox populi—the voice of the people—instigated the regulations that play such a
critical role in protecting workers and the environment. Worker safety and environ-
mental movements [5, 6] are international, with long and interesting histories.
included a 1636 fishermen’s strike on an island near Maine, a 1677 strike by car-men
in New York City, and a 1746 strike by carpenters in Savannah, Georgia.
The United States Congress passed the first federal law governing mine safety in
1891 [9]. The law established ventilation requirements for underground coal mines
and prohibited owners from employing children under 12. Owners had resisted acting
voluntarily, arguing that doing so would put them at a competitive disadvantage.
In 1910, after a decade in which coal mine fatalities exceeded 2000 annually,
Congress established the Bureau of Mines to conduct research and reduce accidents
in the coal mining industry. Regulations were implemented, but violations were
rampant until 1941, when Congress empowered federal inspectors to enter mines.
Over the years, regulations improved, culminating in the Federal Mine Safety Act
of 1977, which was amended by the MINER Act of 2006.
US child labor laws lagged the rest of the world by decades. According to the 1890
U.S. Census, about 1.12 million children under ten years old were engaged in “gain-
ful occupations” out of a total population (including adults) of 63.0 million [10–12].
According to the 1900 US Census, the number had increased to 1.75 million children
out of a total population of 76.2 million. About 16.7% of children under ten were
working, and 18% of children between 10 and 15 were employed full time. Child
labor on family farms was expected, but the increase of 50% in under-10 employ-
ment for wages, mostly in factories, alarmed many Americans. Attempts at reform
were thwarted by the Supreme Court. In 1912, a law established the United States
Children’s Bureau in the Department of Commerce and Labor. In 1916, the Keating-
Owen Act prohibited interstate shipment of goods manufactured or processed by
child labor. The bill passed 337 to 46 in the House and 50 to 12 in the Senate. But
in 1918, in a 5-4 decision, the law was ruled unconstitutional by the Supreme Court.
In 1938—after years of court decisions overturning attempts at labor reform—and
118 years after the first child labor law in England—Congress passed the Fair Labor
Standards Act (FLSA) [13]. The FLSA banned oppressive child labor, set a minimum
wage of 25 cents per hour, and limited the work-week to 44 h. The FLSA survived
Supreme Court challenges, including the United States v. Darby Lumber Co. case in
1941. The FLSA is in force today.
This section describes the predominant gaseous, liquid and solid pollutants generated
by the coal, petroleum, and petrochemical industries.
In response to environmental regulation, the petroleum industry reduced pollution
by the actions listed in Sect. 18.1 under “Air Quality”. The technology behind these
actions is explained below.
In refineries, coking operations and FCC regenerators are the main sources of PM
emissions. Coke-derived PM10 can be reduced by building enclosures around coke-
handling equipment—conveyor belts, storage piles, rail cars, barges, and calciners.
For flue gas from FCC regenerators, many licensors offer scrubbing technology.
ExxonMobil offers wet-gas scrubbing (WGS), which removes particulates, SOx and
NOx [26]. UOP and Shell Global Solutions offer third-stage separator (TSS) tech-
nology, which removes PM in conjunction with flue-gas power recovery [27].
The main sources of air pollution from petroleum refineries are listed in Table 18.1.
Refineries can be significant sources of particulate matter (PM), which can irritate
402 18 Safety and Environment
the respiratory tract. PM is especially harmful when it is associated with sulfur and
nitrogen oxides (SOx and NOx ).
PM from coal is coal dust. PM from refining comes mainly from two sources—de-
layed coking units and the regenerators of fluid catalytic cracking (FCC) units. FCC
regenerators also emit ammonia, which combines with SOx and NOx in the air to
form ammonium sulfates and nitrates. According to the South Coast Air Quality
Management District (AQMD) [28] in Southern California, 1 ton of ammonia can
generate 6 tons of PM10—airborne particulates with particle diameters less than 10
microns. PM2.5 stands for airborne particulates with diameters less than 2.5 microns.
conditions in refinery hydroprocessing units, the nickel on NiMo catalysts can react
with CO to form nickel carbonyl, an exceptionally hazardous gas.
CO from partial combustion in FCC regenerators is converted to CO2 in CO
boilers. Flue gas from other boilers, process heaters, and power plants can also
contain some CO, which can be diminished by the installation of high-efficiency
burners and/or the implementation of advanced process control.
Fugitive emissions (leaks) from storage tanks, sewers, process units, seals, valves,
flanges, and other fittings [29] can contain both CO and volatile organic compounds
(VOC). Floating roofs can be added to open tanks, and tanks that already have a
roof can be fitted with vapor recovery systems. Open grates above sewers can be
replaced with solid covers. Emissions from seals, valves, etc., can be pin-pointed
with portable combustible-gas detectors. Repairs can then be made at convenient
times, e.g., during a maintenance shutdown.
In a carbon adsorption process developed by Lurgi, hot flue gas first goes through
a cyclone or dust collector for particulate removal. The gas is cooled with water
and sent to an adsorption tower packed with activated carbon. The carbon adsorbs
SOx . Water is sprayed into the tower intermittently to remove the adsorbed gas as a
weak aqueous acid. The scrubbed gas goes out the stack. The acid goes to the gas
cooler, where it picks up additional SOx by reacting with incoming flue gas. Cooler
discharge is sold as dilute sulfuric acid.
In the Reinluft carbon adsorption process, the adsorbent is a slowly moving bed
of carbon. The carbon is made from petroleum coke and activated by heating under
vacuum at 1100 °F (593 °C).
Flue-gas scrubbing with catalytic oxidation (Cat-Ox) is an adaptation of the con-
tact sulfuric acid process, modified to give high heat recovery and low pressure drop.
In the Monsanto process, particulates are removed from hot flue gas with a cyclone
separator and an electrostatic precipitator. A fixed-bed converter uses solid vana-
dium pentoxide (V2 O5 ) to catalyze the oxidization of SO2 to SO3 . Effluent from
the converter goes through a series of heat exchangers into a packed-bed adsorption
tower, where it contacts recycled sulfuric acid. The tower overhead goes through an
electrostatic precipitator, which removes traces of acid mist from the scrubbed gas.
Liquid from the tower (sulfuric acid) is cooled and sent to storage. Some of the acid
product is recycled to the absorption tower.
In a flue-gas desulfurization process from Mitsubishi Heavy Industries (MHI),
manganese dioxide (MnO2 ) is the absorption agent. The final product is ammonium
sulfate (NH4 )2 SO4 , which is an excellent fertilizer. MHI claims better than 90%
removal of SOx with this process.
The Wellman-Lord process uses a solution of potassium sulfite (K2 SO3 ) as a
scrubbing agent. K2 SO3 adsorbs SO2 to give potassium bisulfite (KHSO3 ). The
bisulfite solution is cooled to give potassium pyrosulfite (K2 S2 O5 ). This can be
stripped with steam to release SO2 , which is fed to a sulfuric acid plant.
Arguably, SOx transfer additives are the most cost-effective way to lower SOx
emissions from a full-combustion FCC unit. These materials, first developed by
Davison Chemical, react with SOx in the FCC regenerator to form sulfates. When
the sulfated additive circulates to the riser/reactor section, the sulfate is chemically
reduced to H2 S, which is recovered by amine absorption and sent to a sulfur plant.
Sulfur that would have gone up the stack as SOx goes instead to the sulfur plant as
H2 S.
In some units, SOx additives can reduce FCC SOx emissions by more than 70%.
This can have a dramatic effect on the design and/or operation of upstream and
406 18 Safety and Environment
Like CO and SOx , nitric oxide (NO) and nitrogen dioxide (NO2 ) are emitted by fired
heaters, power plants, and FCC regenerators. NOx also damage respiratory tissues
and contribute to acid rain.
Ozone, a nasty component of smog, is generated by reactions between oxygen
and NOx . The reactions are initiated by sunlight. In the troposphere, ozone reacts
with volatile organic hydrocarbons (VOC) to form aldehydes, peroxyacetyl nitrate
(PAN), peroxybenzoyl nitrate (PBN) and a number of other substances. PAN irritates
nasal passages, mucous membranes, and lung tissue. Collectively, these compounds
are called “photochemical smog.” They harm humans, animals, and plants, and they
accelerate the degradation of rubber and construction polymers. In some areas, smog
looks like a brownish cloud just above the horizon. It makes for spectacular sunsets,
but nothing else about it is good.
In refineries and petrochemical plants, NOx is formed in several ways. In high-
temperature fired heaters, NOx is produced by the reaction of nitrogen with oxygen.
In FCC regenerators, NOx is produced from the nitrogen deposited with coke on
spent catalysts. FCC NOx emissions go up when (a) the catalyst contains more com-
bustion promoter, (b) when oxygen in the flue gas goes up, (c) at higher regenerator
temperatures, and (d) at higher feed nitrogen contents. Combustion promoter is a
noble-metal material that accelerates the reaction between CO and O2 to form CO2 .
By removing CO, the promoter inhibits the following reactions:
Dual-alkali flue-gas scrubbing only removes about 20% of the NOx from a typical
flue gas. Therefore, instead of simple scrubbing, chemical reducing agents are used.
In selective catalytic reduction (SCR) processes, anhydrous ammonia is injected into
the flue gas as it passes through a bed of catalyst at 500–950 °F (260–510 °C). The
reaction between NOx and ammonia produces N2 and H2 O.
The MONO-NOx process offered by Huntington Environmental Systems employs
a non-noble metal catalyst. For SOx , NOx and VOC removal, Ducon uses ceramic
honeycomb or plate-type catalysts in which titanium dioxide is the ceramic and the
active coatings are vanadium pentoxide and tungsten trioxide (WO3 ). The working
catalyst temperature ranges from 600 to 800 °F (315–427 °C). For NOx abatement,
Ducon provides complete ammonia injection systems with storage tanks, vaporizers
and injection grids. Either anhydrous or aqueous ammonia can be used.
NOx -removal catalysts are offered by Haldor-Topsøe, KTI, and others. The Ther-
mal DeNOx process offered by ExxonMobil is non-catalytic.
18.4 Pollution Control and Abatement Technology 407
The previous section mentions the harmful effects of ground-level ozone. The strato-
sphere, located about 6–30 miles (10–50 km) above the ground, contains a layer of
ozone that is beneficial, because it protects organisms from harmful ultraviolet-B
(UV-B) solar radiation.
In the 1980s, scientists noticed that the stratospheric ozone layer was getting
thinner [32]. Researchers linked several man-made substances, mostly volatile halo-
genated compounds, to ozone depletion, including carbon tetrachloride (CCl4 ), chlo-
rofluorocarbons (CFCs), halons, methyl bromide, and methyl chloroform. These
chemicals leak from air conditioners, refrigerators, insulating foam, and some indus-
trial processes. Winds carry them through the lower atmosphere into the stratosphere,
where they react with strong solar radiation to release chlorine and bromine atoms.
These atoms initiate chain reactions that consume ozone. Scientists estimate that a
single chlorine atom can destroy 100,000 ozone molecules.
In 1987, twenty-seven countries signed the Montreal Protocol, which recognized
the international consequences of ozone depletion and committed the signers to
limit production of ozone-depleting substances. Today, more than 190 nations have
signed the Protocol, which now calls for total elimination of certain ozone-depleting
chemicals.
The 1998 and 2002 Scientific Assessments of Stratospheric Ozone firmly estab-
lished the link between decreased ozone and increased UV-B radiation. In humans,
UV-B is linked to skin cancer. It also contributes to cataracts and suppression of
the immune system. The effects of UV-B on plant and aquatic ecosystems are not
well understood. However, the growth of certain plants can be slowed by excessive
UV-B. Scientists suggested that marine phytoplankton, which are the foundation of
the ocean food chain, were already under stress from UV-B.
In the United States, production of halons ended in January 1994 [33]. In Jan-
uary 1996, production virtually ceased for several other ozone-depleting chemicals,
including CFCs, CCl4 , and methyl chloroform. January 1, 2010 saw a ban on produc-
tion, import, and use of HCFC-22 and HCFC-142b, except for continuing servicing
needs of existing equipment. In 2015, the production, import, and use of all HCFCs
was banned, except for continuing servicing needs. In 2020, there will be no pro-
duction or imports of HCFC-22, and HCFC-142b, achieving a total reduction of
99.5%.
New products less damaging to the ozone layer have gained popularity. For exam-
ple, computer makers now use ozone-safe solvents to clean circuit boards, and auto-
mobile manufacturers use HFC-134a, an ozone-safe refrigerant, for air conditioners
in new vehicles.
Studies indicate that the Montreal Protocol has been effective. Stratospheric con-
centrations of methyl chloroform are falling, indicating that emissions have been
reduced. Concentrations of other ozone-depleting substances, such as CFCs, are
also decreasing. According to the U.S. EPA [34], the ozone layer has not grown thin-
408 18 Safety and Environment
ner since 1998 and appears to be recovering. Antarctic ozone is expected to return
to pre-1980 levels sometime between 2060 and 2075.
Life, as we know, requires liquid water and nutrients. Macronutrients include com-
pounds containing carbon, hydrogen, sulfur and nitrogen. Arguably, the most funda-
mental food-chain macronutrient is carbon dioxide. Water and carbon dioxide play
key roles in regulating planetary temperatures.
Water is an effective heat buffer, with both a high heat of fusions (freezing and
melting) and a high heat of vaporization (vaporizing and boiling). It takes a lot of
heat to vaporize water completely at its boiling point, and it also takes a lot of heat to
melt ice at water’s freezing point. On earth, water is present in three phases: vapor,
liquid, and ice—mostly liquid. In the atmosphere, water vapor condenses and falls
to the surface as rain, snow, etc. In cold regions, snow collects into ice sheets, which
partially or completely melt during the summer. Rain and melt water collect into
rivers, lakes, and oceans, where evaporation returns water vapor to the atmosphere.
Table 18.2 provides estimates of the global carbon distribution [38]. On per-mole
basis, sedimentary rock contains more than 99.8% of the carbon in the earth crust.
Most of the rest is found in the sea. In oceans and lakes, water serves as a carbon
dioxide buffer. On average, the oceans hold 60 times more CO2 than the atmosphere.
18.4 Pollution Control and Abatement Technology 409
Dissolved in water, CO2 forms carbonic acid (H2 CO3 ), bicarbonate (HCO3 − ) and
carbonate (CO3 2− ). Sea water is slightly alkaline, with a surface pH of 8.2, so it readily
reacts with H2 CO3 . However, rapid exchange with the atmosphere only occurs in the
upper wind-mixed layer, which is about 300 feet (100 m) thick. This layer contains
roughly one atmosphere equivalent of CO2 . CO2 in its various forms is removed from
the sea by foraminifera, coral reefs, and other marine organisms, which produce solid
calcium carbonate, the main component of sea shells.
About 30–50% of the CO2 released into the air stays there. The rest goes into the
hydrosphere and biosphere.
Only 0.0025% of the earth’s carbon is present as atmospheric CO2 . But even
small changes have a dramatic impact on surface temperature balance. The air retains
about 30–50% of the excess CO2 it receives from combustion of fossil fuels, cement
manufacturing, and other human activities. The rest goes into the hydrosphere and
biosphere. An alarming trend is ocean acidification. Coral reefs moderate the impact
of CO2 on ocean pH, but as they so, they slowly dissolve [39].
Atmospheric CO2 concentrations are increasing faster than ever, and most of
the increase is due to industrial growth. According to the Goddard Institute for
Space Studies (GISS), which is administered by the U.S. National Aeronautics and
Space Administration (NASA), the average global temperature has increased 1.4 °C
between 1850 and 2005 [40]. ModelE2 is a GISS application which indicates that
man-made CO2 is responsible for most of the change.
410 18 Safety and Environment
Fig. 18.2 Data from ModelE2, developed by NASA’s GISS group. Shown are forces affecting
global temperatures, 1850–2005. Man-made forces include aerosols (which decrease temperature),
human-generated greenhouse gases, man-made ozone, and land use (deforestation). Natural factors
include periodic changes in the earth’s orbit, solar temperature variations, and volcanic gases (which
include CO2 )
It generally has high salinity, and sometimes contains low levels of radiation. The
wastewater goes to a settling pit, in which entrained oil disengages and floats to
surface. The oil is skimmed off and sold. The skimmed water is trucked to disposal
wells and injected thousands of feet underground for permanent storage. However, if
an incident forces chemicals into an aquifer, responders may be unprepared to take
actions until they analyze the sample. Citizens groups are concerned about ground-
water contamination. The Texas Tribune article cites reports of fracturing causing and
earthquakes and an instance of associated aquifer contamination. Certain aspects of
fracking were exempted from the US Safe Drinking Water Act by the Energy Policy
Act of 2005.
In the downstream, refineries and chemical plants generate contaminated process
water, oily runoff, and sewage. Water is used by just about every process unit, espe-
cially those that require wash water, condensate, stripping steam, caustic, or neutral-
ization acids. Contaminated process water may contain suspended solids, dissolved
salts, water-soluble chemical, phenols, ammonia, sulfides, and other compounds. As
much as possible, waste-water steams are purified and re-used. Present requirements
ensure that the water going out of a processing plant is at least as clean as the water
coming in.
Table 18.3 summarizes the sources and destinations of waste water in the refinery.
Wastewater steams are purified and reused as much as possible. Present requirements
412 18 Safety and Environment
in the US and EU state that the water going out of a process plant must be as clean
as the water coming in.
Primary treatment uses a settling pond to allow most hydrocarbons and suspended
solids to separate from the wastewater. The solids drift to the bottom of the pond,
18.5 Waste Water from Petroleum Production and Processing 413
Inspection ports
Belt Skimmer
Outlet
Oil Film Oil Film Inlet
Sludge Hopper
hydrocarbons are skimmed off the top, and oily sludge is removed. Difficult oil-in-
water emulsions are heated to expedite separation.
Acidic wastewater is neutralized with ammonia, lime, or sodium carbonate. Alka-
line wastewater is treated with sulfuric acid, hydrochloric acid, carbon dioxide-rich
flue gas, or sulfur.
Figure 18.3 a simplified sketch of an API oil-water separator. The large capacity
of these separators slows the flow of wastewater, allowing oil to float to the surface
and sludge to settle out. They are equipped with a series of baffles and a rotating
endless-belt skimmer, which recovers floating oil. Accumulated sludge is removed
through sludge hoppers at the bottom.
A small amount of suspended solids remains in the water after primary treatment.
These are removed by filtration, sedimentation or air flotation. Flocculation agents
may be added to consolidate the solids, making them easier to remove by sedi-
mentation or filtration. Activated sludge, which contains waste-acclimated bacteria,
digests water-soluble organic compounds, in either aerated or anaerobic lagoons.
Steam-stripping is used to remove sulfides and ammonia, and solvent extraction is
used to remove phenols.
414 18 Safety and Environment
Large spills of oil from tankers are not uncommon, but they can cause tremendous
damage to the environment. Small spills come from leaks in tanks or mishaps during
the loading or unloading of trucks, ships, or rail cars.
Oil spills can come from natural seepage, leaky storage tanks, petroleum explo-
ration and production activities, the on-purpose flushing of fuel tanks at sea, and
accidents such as those described in Sect. 18.8.
The cleanup of oil spills includes containment, physical and mechanical removal,
chemical and biological treatment, and natural forces. Land-based spills are easier
to clean than spills onto open water, which are spread quickly by currents and winds.
Several natural forces tend to remove oil spills. These include evaporation, spreading,
emulsification, oxidation, and bacterial decomposition.
Evaporation. A large portion of an oil spill may simply evaporate before other
methods can be used to recover or disperse the oil. Rates of evaporation depend on
the ambient temperature and the nature of the oil.
Spreading. The fact that spilled oil spreads quickly across the surface of water
is a “good news, bad news” story. The good news is that spreading increases rates
of evaporation and air oxidation. The bad news is that the more dispersed the oil
becomes, the harder it is to collect.
As with evaporation, rates of oil-spill dispersion depend upon ambient conditions
and the nature of the oil. Not surprisingly, oil disperses best in fast-moving turbulent
water.
Oxidation. Freshly spilled crude oil has a natural tendency to oxidize in air.
Sunlight and turbulence stimulate the process. Oxidation products include organic
acids, ketones and aldehydes, all of which tend to dissolve in water. As a spill ages,
oxidation slows as “easy” molecules disappear from the mix. Compared to other
natural forces, oxidation plays a minor role in removing oil spills.
18.6 Cleaning Up Oil Spills 415
Emulsification. When crude oil spills at sea, it emulsifies rapidly. Two kinds
of emulsions are formed—oil-in-water and water-in-oil. Oil-in-water emulsions, in
which water is the continuous phase, readily disperse, removing oil from the spill.
However, this kind of emulsion requires the presence of surface-active agents (deter-
gents).
The composition of water-in-oil emulsions varies from 30 to 80% water. These are
extremely stable. After several days, they form “chocolate mousse” emulsions, which
are annoyingly unresponsive to oxidation, adsorption, dispersion, combustion, and
even sinking. The most effective method for mousse emulsions is physical removal.
Mousse contains roughly 80% water, so after a 40–50% loss of light-ends through
evaporation, a spill of 200,000 barrels oil can form 400,000–500,000 barrels of
mousse.
Booms. When oil is spilled on water, floating booms may be used for containment. A
typical boom extends 4 inches (10 cm.) above the surface and 1 foot (30 cm.) below.
Foam-filled booms are lightweight, flexible, and relatively inexpensive. Typically
used for inland and sheltered waters, they are made from polyvinyl chloride (PVC)
or polyurethane. Rectangular floats allow them to be wound onto a reel for storage.
Inflatable booms use less storage space and can be deployed from ships or boats
in open water. Towed booms (Fig. 18.4) are good for preventing dispersion of oil by
winds and currents.
Oil Oil
Water Water
c c Subsurface
Impeller
c
Beach booms are modified for use in shallow water or tidal areas. Water-filled
tubes on the bottom of the boom elevate it above the beach when the water level is
low and allow the boom to float when the water level rises.
Skimmers and pumps. After a spill is contained, skimmers and pumps can pick
up the oil and move it into storage tanks. Weir skimmers are widely used because
they are so simple. Modern designs are self-adjusting and circular so that oil can flow
into the skimmer from any direction. They can be fitted with screw pumps, which
enable them to process many different types of oil, including highly viscous grades.
Cutting knives keep seaweed and trash from fouling the pumps. Screw pumps can
develop high pressure differentials, which gives them higher capacities than other
kinds of pumps.
Another kind of skimmer is an oleophilic (“oil loving”) endless belt. The belt
picks up oil as it passes across the top of the water. As of the belt returns to its
starting point, it is squeezed through a wringer. Oil recovered by the wringer flows
into containers, where it is stored until it can be moved to a land-based processing
facility.
Another method uses a subsurface impeller to create a vortex, similar to the vortex
that forms as water flows out of a bathtub (Fig. 18.5). This vortex funnels water down
through the impeller and creates a bowl of oil in the middle of the vortex. A pump is
used to remove oil from the bowl.
18.6 Cleaning Up Oil Spills 417
18.6.3 Adsorbents
Oil spills can be treated with absorbing substances or chemicals such as gelling
agents, emulsifiers, and dispersants.
When applying adsorbents, it is important to spread them evenly across the oil
and to give them enough time to work. When possible, innocuous substances should
be used. Straw is cheap, and it can absorb between 8 and 30 times its own weight
in oil. When it is saturated, the straw is loaded into boats with rakes or a conveyor
system and transported to land. Oil can be recovered from the straw by passing it
through a wringer.
Synthetic substances may also be used for adsorbing oil spills. Polymers such as
polypropylene, polystyrene, and polyurethane have been successful. Polyurethane
foam is especially good. It can be synthesized onsite easily, even aboard a ship. It
adsorbs oil readily and doesn’t release its load unless it is squeezed. Best of all, a
batch of polyurethane can be used again and again.
Dispersion chemicals act like detergents. One part of the molecule is oil-soluble
while the other is water-soluble. In effect, the oil dissolves in water and diffuses
quickly away from the spill. Dispersants reduce the tendency of oil to cling to partly
immersed solids, such as walls, docks, buoys and boats.
Non-dispersive methods for removing oil spills include gelling, sinking, and burning.
Gelling. Fatty acids and 50% sodium hydroxide can be added to a spill to trigger
a soap-forming reaction. The resulting gel does not disperse. Instead, it remains in
place to block the spread of non-gelled oil.
Oil sinking. Sinking an oil spill keeps it from reaching shorelines, where it can
devastate marine life. Sinking is best used in the open sea. In shallow water near
the coast, it can cause more problems than it solves. Amine-treated sand is the most
common sinking agent. To initiate sinking, a sand/water slurry is sprayed onto the oil
slick through nozzles. The required sand/oil ratio is about 1-to-1. The oil sticks to the
treated sand, which sinks toward the bottom of the sea. According to many experts,
the oil-coated sand remains in place for many years. According to others, it can
damage fragile ecosystems on the ocean floor and/or lead to shellfish contamination.
During a full-scale test 15 miles from the coast of Holland, Shell Oil used amine-
treated sand to sink a 100-ton slick of Kuwaiti crude in less than 45 min. Oil removal
exceeded 95%.
418 18 Safety and Environment
Other materials have been used for oil sinking. These include talc, coal dust,
cyclone-treated fly ash, sulfur-treated cement, and chalk. In general, a 1-to-1 ratio of
sinker weight to oil weight is needed to sink a fresh spill. If weathering takes place,
the density of the oil increases and less sinking agent is needed. It has been estimated
that the cost for sinking is similar to the cost for dispersion. However, in open seas
under high winds and waves, it may be difficult to spread the sinking agent.
Burning. Freshly spilled crude oil in a confined area may be combustible. How-
ever, after several hours, the spill may have thinned due to spreading and the most
volatile components may have evaporated. If so, it may not be possible to ignite the
remaining material. The addition of gasoline or kerosene can restore combustibility.
Burning is not used much anymore. It seldom removes much oil, and it can
generate concentrated, unpredictable pockets of atmospheric CO and SO2 , both of
which are poisonous.
For cleaning oil from beaches, farm machinery and earthmoving equipment have
been used to good effect. In many cases, a layer of straw is spread across the oil.
After a few days, the oil-laden straw is raked onto a conveyor, screened to drop out
sand, and sent to wringer. Recovered oil is trucked away to a refinery. The spent
straw, which still contains some oil, can be blended with coal and burned in a power
plant, or simply incinerated. The separated sand is washed and returned to the beach.
When beach pollution is severe, oily sand is removed with earthmoving equip-
ment. When beach pollution is mild, sand removal may not be needed. Instead,
detergents can be used. They must be applied cautiously to minimize harm to marine
life. Wave action does a great job of mixing detergent into the sand. Usually, the
detergent is applied about one hour before high tide. When the tide comes in, the
washing begins. If high-tide washing is inappropriate, high-pressure hoses can be
employed. Hoses are also effective for cleaning oil off of walls and rocks.
Froth flotation. In the froth-flotation process, oil-soaked sand from a polluted
beach is poured into a vessel, where it is mixed with water and cleaned with a froth
of air bubbles. Aided by chemical or physical pre-treatment, the froth strips oil away
from the sand. Due to its low density and the action of the bubbles, the oil floats to
the top of the vessel, where it is drawn off and sent to a separating chamber, where
entrained water is removed. Tests show that sand containing 5000 ppm of oil can be
cleaned down to 130 ppm, generating effluent water with 165 ppm of oil. Usually,
the cleaned sand is returned directly to the beach.
Hot water cleanup. When milder methods fail to give the desired degree of
cleanup, hot water can mobilize some of the oil still trapped within the polluted
sand. This method is used as a last resort because of the damage it does to inter-tidal
ecosystems.
18.7 Solid Waste 419
Solid wastes from coal and petroleum processing may include the following:
• Coal fines
• Cuttings from oil-well drilling
• Used drilling mud
• Spent catalyst ad catalyst fines
• Acid sludge from alkylation units
• Sludge from the bottom of storage tanks
• Miscellaneous oil-contaminated solids.
Waters that cannot be recycled are cleaned on site, sent to landfills, or transported
to reclamation facilities.
Contaminated solids are produced during the drilling of oil wells, the transporta-
tion of crude oil, and in oil refineries. All oil-contaminated solids are considered
hazardous and must be sent to hazardous-waste landfills. The transportation and dis-
posal of hazardous waste costs an order of magnitude more than the transportation
and disposal of sanitary waste. Thus, there is a huge economic incentive to remove
oil from contaminated solids before they leave a site.
Considerable amounts of solid wastes can be generated during oil drilling. Oil
contaminated cuttings must be cleaned before disposal. Table 18.4 shows the sources
of solid wastes in a modern oil refinery. These data, provided by the American
Petroleum Institute, are based on a “typical” 200,000 barrels-per-day high-conversion
refinery. A plant this size produces about 50,000 tons per year of solid waste and
about 250,000 tons per year of waste water. As discussed above, all waste water must
be purified before it leaves the plant.
10 -
Supercritical
9 - Fluid
8 - Pc = 7.4 MPa
7 - Tc = 31°C
Pressure, MPa
6 -
4 -
3 - Triple Point
2 -
1 -
0 -| | | | | | | | |
-80 -60 -40 -20 0 20 40 60 80
Temperature, °C
18.7.1.2 Sludge
According to RCRA, oil-containing sludge from storage tanks and refinery water
treatment facilities is by definition hazardous, and should be sent to a hazardous land
fill. In most cases, a lot of the sludge can be dissolved with detergents and/or or
solvents (such as hot diesel oil) and blended into crude oil. Alternatively, dissolved
sludge can go to delayed cokers, asphalt plants, carbon black plants, or cement kilns.
In one method, hot water and a chemical are circulated through the tank. On top
of the water, a hydrocarbon such as diesel is added. The density difference between
warm water and the hydrocarbons in the sludge causes the sludge to rise, allowing
the chemical to strip out water and solids. The method recovers good-quality oil,
which can be processed in the refinery, and leaves behinds a relatively clean layer of
solids on the tank bottom.
Tanks associated with slurry oil from FCC units present an interesting challenge.
It can be very expensive to take these tanks out of service, the sludge is loaded with
finely divided FCC catalyst particles, and slurry-oil sludge is difficult to dissolve.
Recently, a process called Petromax has been used to liquefy slurry-oil sludge, allow-
ing it to be pumped out of tanks with ordinary equipment, even while the tanks are
still in use [46].
Some service companies use robots, cutting wands, and other sophisticated
devices to clean tanks completely without sending people inside. These methods
are especially valuable when the tank contents are toxic. One such company is
Petrochemical Services, Inc., which pioneered the use of robots in tank-cleaning
operations.
Blending mobilized sludge into crude oil is limited by specifications on basic
sediment and water (BS&W). This seems equivalent to moving waste from one
place to another, but it really isn’t. The dissolution of tank sludge separates useful
oil from inorganic solids (sand, clay, salts, and metal oxides) and refractory organics
(asphaltenes, long-chain waxy paraffins, kerogen, and coke).
422 18 Safety and Environment
Many refinery catalysts are regenerated several times. The regeneration of FCC
catalysts occurs in the (aptly named) regenerator. Catalysts from fixed-bed units can
be regenerated in place, but usually they are sent off-site to a facility that specializes
in catalyst regeneration.
Eventually, even the hardiest catalysts reach the end of their useful life. When this
happens, they are sent to a metal reclaimer, which recovers saleable products such
as alumina, silica, MoO3 , V2 O5 , nickel metal, and various forms of cobalt.
The reclamation company makes money on both ends of the plant—from the
refiner who must dispose of the catalyst, and from the customer who buys reclaimed
products.
Coal mine explosions and maladies such as black-lung disease have plagued the
industry since the start of the Industrial Revolution. In 1906, in a mine explosion in
Courrières, France, 1099 miners died, including children. Courrieres was the worst
mine accident in Europe. In April 1942, in a coal-mine explosion in Benxi, Liaoning,
China, 1549 workers died in the worst coal mine accident ever.
Mine explosions are caused by methane and/or coal dust. “Damp” is a catch-all
term for gases other than air in coal mines. In addition to firedamp (methane), there is
blackdamp (carbon dioxide), stinkdamp (hydrogen sulphide), and afterdamp (carbon
monoxide). Firedamp is explosive in air at concentrations between 4 and 16%.
In May 1812, Sir Humphry Davy developed the Davy lamp, in which a low flame
was surrounded by iron gauze. Due to the gauze, the lamp did not ignite external
methane, but methane could pass into the lamp and burn safely above the flame. At
first, the Davy lamp improved safety by eliminating open-flame lighting. But it led to
the mining of areas that had previously been closed for safety reasons, sometimes with
disastrous results [47]. Black lung disease is actually a set of maladies associated with
coal dust [48]. It was not well-understood until the 1950s. In the Federal Coal Mine
Health and Safety Act of 1969, the US Congress set limits on coal dust exposure and
created the Black Lung Disability Trust. Mining companies agreed to a clause which
guaranteed compensation for workers who had spent 10 years doing mine work,
coupled with X-ray or autopsy evidence of lung damage. Financed by a federal tax
on coal, by 2009 the Trust had distributed over $44 billion in benefits to disable
miners or their widows.
The Lakeview gusher remains the largest accidental oil leak or spill on land in
history. It released 1.2 million tons of crude in 544 days. Drilling started in January
1909. Initially, the only product from the well was natural gas. On March 14, 1910,
pressurized oil blew through the well casing above the bit. At the time, blowout
prevention technology didn’t exist, so the well gushed unabated. The initial flow
rate was 18,800 barrels per day. The peak flow rate was 90,000 barrels per day. The
blowout created a river of crude, which crews tried to contain with sand bag dams
and dikes. About half of the oil was saved. The remainder seeped into the ground or
evaporated.
424 18 Safety and Environment
On April 16, 1947, the SS Grandcamp exploded in port at Texas City, Texas [50].
The ship was loaded the 7700 tons of ammonium nitrate. The initial blast, together
with fires and explosions on other ships and nearby oil-storage facilities, killed 581
people and injured up to 8000. The force of the blast was equivalent to 3.2 kilotons
of TNT. At about 8:00, a fire was spotted in the cargo hold of the ship. For roughly
an hour, attempts to extinguish the fire failed as a red glow returned after each effort
to douse the fire. The cause of the initial fire remains unknown. The event drew
attention to (a) the dangers of storing massive amounts of hazardous materials and
(b) storing hazardous materials in polluted areas.
In December 1952, thick fog rolled across many parts of the British Isles. In the
Thames River Valley, the fog mixed with smoke, soot and sulfur dioxide from coal-
burning homes and factories, turning the air over London into a dense yellow mass.
The smog was so thick that buses could not run without guides walking ahead of
them carrying lanterns. Fog stayed put for several days, during which the city’s
hospitals filled to over-flowing. According to the Parliamentary Office of Science
and Technology [51], more than 3000 people died that month because of the polluted
air. Similar, less-severe episodes occurred in 1956, 1957, and 1962. The 1956 event
killed more than 1000 people.
London’s deadly smog was caused by “usual-practice” pollution. Due to the
widespread use of cheap, high-sulfur coal, the air in London had been bad for decades,
but post-war growth made it worse than ever. In response to the incidents, Parliament
passed Clean Air Acts in 1956 and 1962, prohibiting the use of high-sulfur fuels in
critical areas.
On March 16, 1978, the supertanker Amoco Cadiz was three miles off the coast of
Brittany, France when its steering mechanism failed. The ship ran aground on the
Portsall Rocks.
For two weeks, severe weather restricted cleanup efforts. The wreck broke up
completely before any remaining oil could be pumped out, so the entire cargo—more
than 1.6 million barrels of Arabian and Iranian crude oil—spilled into the sea.
The resulting slick was 18 miles wide and 80 miles long. It polluted 200 miles of
coastline, including the beaches of 76 Breton communities, harbors and habitats for
18.8 Significant Accidents and Near-Misses 425
marine life. On several beaches, oil penetrated the sand to a depth of 20 inches. Piers
and slips in small harbors were covered with oil. Other polluted areas included the
pink granite rock beaches of Tregastel and Perros-Guirrec, and the popular bathing
beaches at Plougasnou. The oil persisted for only a few weeks along exposed rocky
shores, but in the areas protected from wave action, the oil remained as an asphalt
crust for several years.
At the time, the Amoco Cadiz incident caused more loss of marine life than any
other oil spill. Cleanup activities on rocky shores, such as pressure-washing, also
caused harm. Two weeks after the accident, millions of dead mollusks, sea urchins,
and other bottom-dwelling organisms washed ashore. Nearly 20,000 dead birds were
recovered. About 9000 tons of oysters died. Fish developed skin ulcerations and
tumors.
Years later, echinoderms and small crustaceans had disappeared from many areas,
but other species had recovered. Even today, evidence of oiled beach sediments can
be seen in sheltered areas, and layers of sub-surface oil remain under many impacted
beaches.
A 2.5 mile permanent boom protected the Bay of Morlaix. Although it required
constant monitoring, the boom functioned well because the bay was protected from
severe weather and the brunt of the oil slick. In other areas, booms were largely inef-
fective due to strong currents, and also because they were not designed to handle such
enormous amounts of oil. Skimmers were used in harbors and other protected areas.
Vacuum trucks removed oil from piers and boat slips where seaweed was especially
thick. “Honey wagons”—vacuum tanks designed to handle liquefied manure—were
used to collect emulsified oil along the coast.
Oil-laden seaweed was removed from the beaches with rakes and front-end load-
ers. Farm equipment was used to plow and harrow the sand, making it more sus-
ceptible to wave and bacterial action. Prior to harrowing, chemical fertilizers and
oleophilic bacteria were applied to the sand.
At first, authorities decided against using dispersants in sensitive areas and along
the coastal fringe. Meanwhile, the spill formed a highly stable water-in-oil emul-
sion (“chocolate mousse”). On the open sea, the French Navy applied both dilute
and concentrated dispersants, but good dispersion was hard to achieve because in
some places the mousse emulsion was several centimeters thick. If dispersants had
been dropped from the air at the source of the spill—in days instead of weeks—the
formation of mousse emulsion might have been prevented.
About 650 metric tons of chalk was applied in an effort to sink the oil. But after
one month at sea, the oil was so viscous that the chalk just sat of top of it. Rubber
powder made from ground-up tires was applied to absorb the oil. The French Navy
used water hoses to spread most of the powder. Some was applied manually from
small fishing boats. Because it stayed on top of the oil, the rubber powder had little
effect; wave action wasn’t strong enough to mix it into the oil, most of which was
trapped inside the chocolate mousse emulsion.
During the third and fourth months of the cleanup, high-pressure hot water (fresh
water at 2000 psi, 80–140 °C) was very effective in cleaning oil from rocky shores.
A small amount of dispersant was applied to prevent oiling of clean rocks during
426 18 Safety and Environment
the next high tide. The mouths of several rivers contained oyster beds and marshes
that required manual cleaning. Soft mud river banks were cleaned with low-pressure
water. To improve oil-collection efficiency, a sorbent was mixed with water and
poured in front of the wash nozzles. The oil was collected downstream by a local
invention called an “Egmolap.” This device was good at collecting floating material
in sheltered areas.
The literature contains many summaries of the causes and lingering impact the Cher-
nobyl nuclear disaster. The International Nuclear Safety Advisory Group (INSAG)
published a thorough report in 1986 and followed it with a revision in 1992 [53].
The INSAG documents remain the most generally accepted. The oft-vilified Anatoly
Dyatlov, who was the Deputy Chief Engineer at the plant when the incident occurred,
challenged the INSAG report. An English version of his perspective appeared in 1995
[54]. There are numerous summaries and timelines on the Internet, but they often
disagree with each other and the facts in the INSAG document. A factual summary
of health effects was published by the World Nuclear Association in 2009 [55]. Other
articles on health effects are outlandish. For this chapter, choices had to be made.
Please accept apologies for any errors.
A “SCRAM” is an emergency shutdown of a nuclear reactor. According to lore, it
is an acronym for “safety control rod axe man.” Supposedly, Enrico Fermi coined the
term for the first nuclear reactor at the University of Chicago where graphite control
rods were raised and lowered into the reactor core with a rope on a pulley. The axe
man stood ready to cut the rope in an emergency, dropping all control rods all the
way down, shutting down the reactor.
After a SCRAM at the Chernobyl nuclear power plant near Pripyat, Ukraine, the
cooling water pumps stopped due to lost electrical power. Emergency power was
supplied by diesel generators, but starting the diesels took almost a minute. During
the delay, residual heat remained in the reactor until the cooling pumps were running
again.
To Anatoly Dyatlov, this delay was unacceptable. He thought of a solution. Right
after a SCRAM, he knew, the plant’s power turbines kept spinning for several minutes
due to their massive inertia. Dyatlov calculated that, as the turbines slowed down,
they could produce enough power to run the water pumps until the diesels were up
and running. He designed a complex network of switches to keep the current steady
as the turbines lost momentum.
Dyatlov talked to the Chief Engineer of the plant, who gave him permission to
run his experiment. The test began on April 25, 1986, as Unit 4 was shutting down
for long-delayed maintenance. With Dyatlov’s approval, the operators decided to run
the test manually instead of using the unit’s “unimaginative automatics.”
At 2:00 PM, in violation of one of the most important safety regulations, the
operators switched off the emergency cooling system. Just after midnight on April
26th, they switched off the reactor’s power-density controls.
Under manual operation, the reactor became unstable. At 1:07 AM, the power
output suddenly dropped to 0.03 gigawatts (GW), far less than the specified minimum
of 0.70 GW. To generate more power, the operators raised the graphite control rods to
the maximum height allowed by regulations. Dyatlov ordered them to raise the rods
further. When the operators balked, Dyatlov insisted, and the operators complied.
At 1:23 AM, the power output seemed to be stable at 0.2 GW. The operators then
violated THE most important safety regulation by disabling the emergency SCRAM,
428 18 Safety and Environment
an automatic interlock designed to stop the reactor whenever the neutron flux exceeds
a safe limit. (In modern nuclear power plants, it is physically impossible to disable
this control.)
Now the experiment could begin. To see how long the electricity-generating tur-
bine would spin without a supply of steam, they closed the valve that channeled
steam from the reactor to the turbine. Steam that should have been taking heat out
of the reactor was now trapped inside.
In less than 45 s, the reactor started to melt. Super-hot pellets of uranium-oxide fuel
ruptured their zirconium-alloy containers, coming into direct contact with cooling
water. The water flashed into steam, causing the first of two explosions that blew the
top off the reactor. The second blast was caused by a H2 -CO-air explosion. The H2
and CO were generated by reactions of zirconium and graphite with super-heated
steam:
About 15 tons of radioactive material from the reactor core rocketed into the
atmosphere, where it spread across Europe. More than 36 h after the accident, plant
personnel told local officials about the accident. About 14 h after that—50 h after the
accident—radiation from the explosion was detected by technicians at the Forsmark
nuclear power plant in Sweden. That measurement was the first notification to the
world outside the Soviet Union that something had happened in Pripyat.
At the site, more than 30 fires broke out, including an intensely hot graphite
fire that burned for 14 days. About 250 people fought the various fires. During 1800
helicopter sorties, pilots dropped 5000 tons of lead, clay, dolomite and boron onto the
reactor. Near term, the explosion and high-level radiation killed 31 people, including
operators, fire-fighters, and helicopter pilots. Thousands more died later.
During a trial that ended in August 1987, the Chernobyl plant director (Brukhanov)
was convicted and sentenced to 10 years in a labor camp. The chief engineer (Fomin),
and Dyatlov received shorter sentences. The two operators were acquitted, but both
of them died soon afterwards from radiation poisoning. More than 60 other workers
were fired or demoted.
In June 1987, 14 months after the disaster, some 27 villages within the restricted
zone were still heavily contaminated, because the cleanup operation had stopped.
Nearby cities and towns were reporting dramatic rises in thyroid diseases, anemia,
and cancer. Hardest hit were children. Frequently, calves were born without heads,
eyes, or limbs.
According to government figures, more than 4000 Ukrainians who took part in
the clean-up had died, and 70,000 were disabled by radiation. About 3.4 million of
18.8 Significant Accidents and Near-Misses 429
Ukraine’s 50 million people, including some 1.26 million children, were affected
by Chernobyl [56]. According to Gernadij Grushevoi, co-founder of the Foundation
for the Children of Chernobyl, the long-term danger is even worse in Belorussia.
In 1992, in testimony at the World Uranium Hearing [57], he said, “… 70% of the
radioactive stuff thrown up by the explosion at Chernobyl landed on White Russian
territory. There is not a centimeter of White Russia where radioactive cesium cannot
be found.”
18.8.7.2 The Rest of the Plant Kept Running for Nine Years
The three remaining nuclear reactors (Units 1, 2 and 3) continued to run. Despite
ambient radiation levels 9 times higher than widely accepted limits, workers lived in
newly built colonies inside the “dead zone.”
In 1991, Unit 2 was damaged beyond repair by a fire in the turbine room. That
left Units 1 and 3, which kept going for the next nine years. They kept going because
the power was needed, and no funds were available to build a replacement.
On December 14, 2000, the plant was shut down. The shut-down was expedited by
the European Commission, which approved a US$585 million loan to help Ukraine
build two new reactors, and by the European Bank for Reconstruction and Develop-
ment, which provided US$215 million.
On November 1, 1986, a fire broke out in a riverside warehouse at the Sandoz chem-
ical plant in Schweizerhalle, Switzerland. While extinguishing the flames, firemen
sprayed water over exploding drums of chemicals, washing as much as 30 tonnes
of pesticides, chemical dyes, and fungicides into the Rhine River. Up to 100 miles
(160 km) downstream from Schweizerhalle, the Rhine was sterilized. All told, more
than 500,000 eels and fish were killed. More than 50 million people in France, Ger-
many, and The Netherlands endured drinking water alerts.
Afterwards, while checking the Rhine for chemicals as it rolled through Germany,
officials discovered high levels of a herbicide (Atrazine) that wasn’t on the list pro-
vided by Sandoz. Eventually, Ciba-Geigy admitted that it had spilled Atrazine into
the river just a day before the Sandoz fire. As monitoring continued, more chem-
icals were discovered, alerting German authorities to the fact that many different
companies were secretly and knowingly discharging dangerous chemicals into the
river. BASF admitted to spilling more than a tonne of herbicide, Hoechst discov-
430 18 Safety and Environment
ered a chlorobenzene leak, and Lonza confessed to losing 2000 gallons (4500 L) of
chemicals.
18.8.8.2 Recovery
In response to the Sandoz disaster, companies all along the Rhine joined the Rhine
Action Program for Ecological Rehabilitation, agreeing to cut the discharge of haz-
ardous pollutants in half by 1995. Although many experts thought the target could
never be reached, samples showed that from 1985 to 1992, mercury in the river at the
German town of Bimmen-Lobith, near the Dutch border, fell from 6.0 to 3.2 tonnes,
cadmium from 9.0 to 5.9 tonnes, zinc from 3600 to 1900 tonnes, and polychlorinated
biphenyls (PCBs) from 390 to 90 kilograms.
In December 1990, for the first time in 3 years, a large Atlantic salmon was fished
from the Sieg, a tributary of the Rhine in West-Central Germany. The catch proved
what officials had hoped: If you clean up your mess and clear the way, someday
life will return. The event gave impetus to the Salmon 2000 project, and further
success followed. In 1994, researchers found recently hatched salmon in the Sieg,
and in 1996 a salmon was hooked near Baden-Baden. In 1998, encouraged by the
success of the Rhine Action Program, targets were set to designate a large protected
ecosystem from streams in the Jura Mountains, the Alps, the Rhine mountains, the
Rhineland-Palatinate, the Black Forest, and the Vosges to the mouth of the Rhine in
The Netherlands. Meanwhile, not all of the Rhine’s pollution problems have been
solved. One of the most serious is a huge basin in the Netherlands, into which toxin-
laden mud dredged from the Port of Rotterdam has been dumped since the 1970s.
Contamination levels are falling, but several toxins are very stubborn.
All along the Rhine, the main source of remaining pollution comes from farm
fertilizers, which seep into the river every time it rains.
On March 23, 1989, the 987-foot supertanker Exxon Valdez left port carrying more
than 1.2 million barrels of Alaskan North Slope crude. The ship was headed south
toward refineries in Benicia and Long Beach, California. At 10:53 PM, it cleared the
Valdez Narrows and headed for Prince William Sound in the Gulf of Alaska. To avoid
some small icebergs, Captain Joseph Hazelwood asked for and received permission to
move to the northbound shipping lane. At 11:50 PM, just before retiring to his cabin,
the captain gave control of the ship to the third mate, Gregory Cousins, instructing
him to steer the vessel back into the southbound lane after it passed Busby Island.
Cousins did tell the helmsman to steer to the right, but the vessel didn’t turn sharply
enough. At 12:04 AM, it ran aground on Bligh Reef. It still isn’t known whether
Cousins gave the order too late, whether the helmsman didn’t follow instructions
properly, or if something went wrong with the steering system.
18.8 Significant Accidents and Near-Misses 431
Captain Hazelwood returned to the bridge, where he struggled to hold the tanker
against the rocks. This slowed the rate of oil leakage. He contacted the Coast Guard
and the Alyeska Pipeline Service Company. The latter dispatched containment and
skimming equipment. According to the official emergency plan, this equipment was
supposed to arrive at a spill within 5 h. In fact, it arrived in 13 h—eight hours late.
The Exxon Baton Rouge was sent to off-load the un-spilled cargo and to stabilize
the Valdez by pumping sea water into its ballast tanks. The oil transfer took several
days. By the time it was finished, more than 250,000 barrels of oil had spilled into
the Sound. Eventually, 33,000 birds and 1000 otters died because of the spill.
Eleven thousand workers treated 1200 miles (1900 km) of shoreline around Prince
William Sound and the Gulf of Alaska, using 82 aircraft, 1400 vessels, and 80 miles
(128 km) of oil-containing booms.
In response to the disaster, the U.S. Congress passed the Oil Pollution Act of
1990. The Act streamlined and strengthened the ability of the U.S. Environmental
Protection Agency (EPA) to prevent and react to catastrophic oil spills. A trust fund,
financed by a tax on oil, was established to pay for cleaning up spills when the
responsible party cannot afford to do so. The Act requires oil storage facilities and
vessels to submit plans to the Federal government, telling how they intend to respond
to large oil discharges. EPA published regulations for above-ground storage facilities,
and the U.S. Coast Guard published regulations for oil tankers. The Act also requires
the development of area contingency plans to prepare for oil spills on a regional
scale.
Captain Hazelwood was tried and convicted of illegally discharging oil, fined
US$50,000, and sentenced to 1000 h of community service. Exxon spent US$2.2
billion to clean up the spill, continuing the effort until 1992, when both the State of
Alaska and the U.S. Coast Guard declared the cleanup complete. The company also
paid about US$1 billion for settlements and compensation.
On July 10, 2004, USA Today reported: “Not one drop of crude oil spilled into
Prince William Sound from oil tankers in 2003—the first spill-free year since the
ships started carrying crude from the trans-Alaska pipeline terminal in 1977.”
18.8.10 Gulf War: Intentional Oil Spill, Oil Well Fires (1991)
On January 25–27, 1991, during the occupation of Kuwait, Iraqis pumped 4–6 million
barrels of oil from stations at Mina Al-Ahmadi into the Arabian Gulf. They did so to
preempt an expected beach invasion by allied troops. The size of the spill was 16–25
times greater than the Exxon Valdez spill. On January 27, allied bombers stopped the
spill by destroying the pumping stations.
Ad Daffi Bay and Abu Ali Island experienced the greatest pollution. The spill
damaged sensitive mangrove swamps and shrimp grounds. Marine birds, such as
432 18 Safety and Environment
cormorants, grebes, and auks, were killed when their plumage was coated with oil.
The beaches around the shoreline were covered with oil and tar balls.
Despite the ongoing war, the clean-up of the oil spill proceeded rapidly. Kuwaiti
crude is rich in light ends, and water in the Arabian Gulf water is relatively warm.
For these reasons, about half of the spilled oil evaporated, leaving behind a thick
emulsion which eventually solidified and sank to the bottom of the sea. Another 1.5
million barrels were recovered by skimming. Operators of sea-water cooled factories
and desalination plants were concerned that the oil might foul their intake systems.
To prevent this, protective booms that extended three feet (1 m) below the surface
were installed around intakes in Bahrain, Iran, Qatar, Saudi Arabia, and the United
Arab Emirates.
On February 23–27, 1991, retreating Iraqi soldiers damaged three large refineries
and blew up 732 Kuwaiti oil wells, starting fires on 650 of them. Up to 6 million
barrels per day were lost between February 23 and November 8, 1991. Crews from
34 countries assembled to fight the oil-well fires. Initially, experts said the fires
would rage for several years. But due to the development of innovative fire-fighting
technology, the job took less than 8 months.
The oil-well fires burned more than 600 million barrels, enough to supply the
United States for more than a month. The fire-fighting effort cost US$1.5 bil-
lion. Rebuilding Kuwait’s refineries cost another US$5 billion. In all, Kuwait spent
between US$30 and US$50 billion to recover from the Iraqi invasion.
On January 21, 1997 a temperature excursion led to the rupture of a reactor outlet
pipe at the Tosco Avon Refinery in Martinez, California. The consequent release of
hydrogen and hydrocarbons ignited on contact with air, causing an explosion and fire.
One worker was killed and 46 others were injured. There were no reported injuries
to the public.
18.8.11.1 Context
Fixed-bed catalytic hydrocracking converts vacuum gas oil, and other petroleum
fractions with similar boiling ranges, into diesel, jet fuel, gasoline and other products
[61]. Commercial fixed-bed units employ two kinds of catalyst—hydrotreating and
hydrocracking – which usually are loaded into separate reactors.
18.8 Significant Accidents and Near-Misses 433
Fig. 18.8 Arrangement of reactors in the Tosco Avon hydrocracker (numbers in the reactor are bed
numbers)
Figure 18.8 shows the reactor layout for the Tosco Avon hydrocracker. The unit
has 3 parallel Stage 1 hydrotreating reactors with 6 beds each, and three parallel
Stage 2 cracking reactors with 5 beds each. The effluents from the Stage 1 reactors
are combined and sent to a stripper. The stripper removes light gases, including
methane, ethane, propane, and hydrogen sulfide. The stripper liquids go to three
parallel Stage 2 hydrocracking reactors.
On January 19, two days before the fatal event, operators managed to defeat a tem-
perature excursion in Reactor 1 (R1), without depressuring, by aggressively manip-
ulating quench flow. The outlet pipe temperature indicator (TI) exceeded 482 °C
(900 °F) versus the metallurgical limit of 465 °C, and the center TI in Reactor 1
Bed 4 (R1B4) reached 537 °C (998 °F) versus a specified maximum of 440 °C.
During the excursion, an operator went outside to get TI readings from a field panel
underneath the outlet pipe. The panel TIs could measure higher temperatures than
the control-room readings.
Table 18.5 shows a timeline for the accident on January 21. The problem began
several hours earlier when one of the Stage 1 hydrotreating reactors was brought
down to repair a leak. The total unit feed rate was kept the same. This decreased
the residence time of oil in the remaining two Stage 1 reactors by 50%, decreasing
18.8 Significant Accidents and Near-Misses 435
Fig. 18.9 Inhibition of hydrocracking by ammonia and organic nitrogen, where K(organic N) =
7 × K(NH3 ) . The trends are a function of percentage of inhibited catalyst sites at steady state, which
is related to, but not the same as, the concentrations of ammonia or nitrogen in the feed
Table 18.5 Abbreviated timeline with comments for the Tosco Avon incident
4:50 AM
One of three Stage 1 hydrotreating reactors was brought down to repair a leak. The total unit
feed rate was kept the same, which reduced, the extent of HDN
10:00 AM
Nitrogen slip had risen from <15 to 352 ppmw
11:30 AM
Due to the high nitrogen slip and resulting inhibition of Stage 2 cracking activity, there was no
temperature rise in one cracking reactor and very little rise in the other reactors. No light
products were being formed. Stage 1 temperatures were raised to decrease nitrogen slip
1:10 PM
Stage 2 hydrocracking reactor temperatures were raised to “burn off” inhibitory nitrogen. This
was a horrendous mistake. Treating and cracking temperatures must never be raised at the same
time. Figure 18.4 shows how very sensitive activity is to organic nitrogen
5:38 PM
Nitrogen slip from Stage 1 result was reported to be 47 ppmw. The actual nitrogen slip when the
result was received was considerably lower than 47 ppmw. Since 10:00 AM, the nitrogen slip
had fallen 300 ppmw in 7½ hours, an average of 40 ppmw per hour. If the lag time between
taking a sample and getting results was 1 h as usual, the true nitrogen slip was <10 ppmw at 5:38
PM and probably close to zero at 7:34 PM
7:34 PM
In Reactor 3, Bed 4 (R3B4), one TI jumped 108 °C in 40 s, reaching 439, 10 °C higher than
allowed by procedure but not higher than in previous events
7:34:20 PM
Quench to R3B5 opened 100%. Data logger temperatures cycled between −18, 649, and 343 °C.
In well-run units, the associated uncertainty would have compelled someone to bring the unit
down immediately (if it had not been depressured previously due to high temperature)
Sometime before 7:37 PM
An operator went outside to verify temperatures at the field panel
7:37–7:39 PM
All R3B5 outlet TIs were >415 °C. One reached 679 °C
After 7:40 PM
An operator phoned an instrument technician to work on the fluctuating data logger
7:41:20 PM—seven minutes after the first temperature jump
Explosion
The operator at the field panel died
A board operator activated the EDS
18.8 Significant Accidents and Near-Misses 437
Fig. 18.10 Reactor 3 temperatures during the TOSCO Avon incident. Temperature indicator Pt-
133C2 jumped by 104 °C (186 °F) in 20 s, reaching 439 °C (823 °F). The excursion cascaded
through the quench section into the next catalyst bed, causing a jump of 223 °F (124 °C) in 40 s at
Pt-140C3, reaching 460 °C (860 °F)
Figure 18.10 shows Reactor 3 temperatures during the Tosco temperature excur-
sions. At 7:34 PM, a temperature excursion struck R3B4 and cascaded into R3B5,
peaking at 439 °C (823 °F). According to operating guidelines from the technology
licensor, the unit should have been depressured when any bed TI reading reached
427 °C (800 °F). But instead of depressuring, operators made their most dangerous
decision: To keep the unit running. To do what they had done two days before and
several times before that. They tried to control the excursion by cutting heater firing
and manipulating quench gas. As before, an operator went outside to get readings
from the field panel beneath the reactor outlet pipe.
Operators said they failed to act on January 21, because they didn’t know whether
or not an excursion was occurring. They had to reconcile temperature readings from
438 18 Safety and Environment
three different sets of displays. Two sets were in the control room, and the other,
as mentioned, was in the outside field panel. One control-room system was the data
logger, which included most temperature readings. After the initial excursion, the
data-logger readings started to fluctuate wildly, leading to further confusion.
In other refineries, uncertainty is not an excuse for unsafe behavior. Operators are
told: When in doubt, assume the worse; bring the unit down.
Analysis of the ruptured pipe showed that the failure did not occur at weld, elbow
or reducer. The pipe simply got too hot, reaching 1700 °F (927 °F). At the point
of failure, the circumference of the 10-inch (245 mm) diameter pipe expanded by
5 inches (120 mm) before it failed. The wall thickness at the rupture was 0.3–0.4
(7–10 mm), about 50% lower than the thickness of the rest of the pipe.
Root causes for the Tosco Avon accident include the following:
• Management directives. On July 23, 1992, the unit was depressured at low rate to
stop an excursion. A grass fire started at the flare. A worker told the author (PRR),
in a private conversation, that after this event, 1992, upper management strongly
discouraged use of depressuring. A fire might spread into the nearby grassy hills
and the John Muir Historic Site, creating a public relations nightmare. Fixing the
flare would have required investment, which in those days at Tosco was not an
option.
• Confusion. Operators used three different instrumentation systems to obtain tem-
perature data. Not all the data were immediately accessible, and the data often
disagreed. Monitoring points capable of reading the highest temperatures were
underneath the reactors and not connected to the control room. (In the same pri-
vate conversation with the author of this chapter, a worker said that operations
had lobbied hard to connect field-panel measurement to the control room, but
management declined the requests to save money.)
• Inadequate supervision and process training. Supervisory management and
operator training were inadequate. Operators and engineers did not understand
that default values of zero on the data logger might mean extremely high tempera-
tures. They did not understand that a decrease in makeup hydrogen flow indicated
an extreme temperature excursion. They did not understand the temperature sen-
sitivity of cracking catalyst activity at low N-slip.
• Misunderstanding of sampling dynamics. Operators and engineers did not com-
prehend the dynamics of sampling, i.e., the lag time between taking a sample,
getting results, and what might have happened to the process in the meanwhile.
18.8 Significant Accidents and Near-Misses 439
On March 23, 2005, an explosion in an isomerization unit at the BP Texas City, Texas
refinery [62], along with related explosions and fires, killed 15 people and injured
another 180. A shelter-in-place order was issued, requiring 43,000 people to remain
indoors. Houses were damaged as far away as three-quarters of a mile. Resulting
financial losses exceeded $1.5 billion.
18.8.12.1 Context
The accident occurred during the restart of the isomerization unit, after maintenance
on the raffinate splitter. Before the restart, there were several unresolved safety-
critical issues, including a faulty pressure control valve, a defective high-level alarm
in the splitter, a defective (opaque) sight glass used for level measurement, and an
un-calibrated level transmitter. During the morning of March 23, operators noticed
that the heavy raffinate storage tanks were nearly full. Consequently, the night-shift
supervisor agreed that the startup should not continue until some liquid was with-
drawn from the tanks. But the day-shift supervisor arrived late and did not receive
a full status report, including information about the raffinate tank level. Under his
direction, the startup resumed at 9:30 AM. To reduce level in the raffinate splitter,
the level control valve was opened, sending liquid to the already-nearly-full heavy
storage tank. At about 10:00 AM, the level control valve was shut. As the startup
proceeded, more raffinate entered the tower. Without a way to withdraw product, the
tower started filling. The un-calibrated level transmitter indicated that the level was
below 100%. The opaque sight glass was useless, so it wasn’t possible to confirm
the level visually.
As the startup proceeded, burners in the furnace were started to heat the inflowing
raffinate. The procedure specified that the temperature of return flow to the tower
should be raised to 135 °C (275 °F) at 10 °C (18 °F) per hour, but this time the
procedure was altered. The return flow temperature reached 153 °C (307 °F) with
an increase rate of 23 °C (41 °F) per hour. The defective transmitter was indicating
a high (but still-safe) level of 93%. But due to the closed control valve, the true
level was higher. Heat was increasing, causing expansion of the raffinate. Pressure
started to build, but the operations team thought the increase was due primarily to
overheating, which was a known start-up issue. The pressure was relieved and the
furnaces were turned down.
At 12:42 PM, the level transmitter showed 78%. But the true level was higher,
and due to heat from heat exchange, the liquid was expanding. Liquid overflowed
into the pressure-relief system. At 1:13 PM, all three pressure relief valves opened,
sending liquid to the blow-down drum. An operator opened the level control valve
18.8 Significant Accidents and Near-Misses 441
and shut down the furnace completely. But he did not stop the inflow of raffinate. In
the blow-down drum, a “geyser” of hot hydrocarbon liquid vented directly into the
air, then flowed down the outside of the drum, where it hovered above the ground and
drifted away from the unit. It soon reached an idling pick-up truck about 9 m (30 feet)
away. Investigators concluded that the truck must have ignited the hydrocarbons. The
15 people who died were in or around a nearby row of contractor trailers. I (PRR)
was scheduled to be participating in a hydrocracker startup in the refinery on day the
explosion. The startup was delayed. Otherwise, I might have been in one of those
trailers.
According to the investigation report, the Texas City disaster was caused by organi-
zational and safety deficiencies at all levels of the BP Corporation. Warning signs of
a possible disaster were present for several years, and reported to officials by employ-
ees. But company officials did nothing. The extent of the safety culture deficiencies
was further revealed when the refinery experienced two additional serious incidents
just a few months after the March 2005 disaster. In one, a pipe failure caused a
reported $30 million in damage. The other resulted in a $2 million property loss.
The report cited the following local deficiencies at the Texas City refinery:
• A work environment that encouraged operations personnel to deviate from proce-
dure
• Lack of a BP policy or emphasis on effective communication for shift change and
hazardous operations (such as during unit startup)
• Malfunctioning instrumentation that did not alert operators to the actual conditions
of the unit
• A poorly designed computerized control system that hindered the ability of oper-
ations personnel to determine if the tower was overfilling
• Ineffective supervisory oversight and technical assistance during unit startup
• Insufficient staffing to handle board operator workload during the high-risk time
of unit startup
• Chronic worker fatigue, due to the lack of a human fatigue-prevention policy
• Inadequate operator training for abnormal and startup conditions
• Failure to establish effective safe operating limits.
The industry learned from the BP Texas City disaster, but not everyone learned, and
some who might have learned forgot: similar behavior occurred during the following
event.
442 18 Safety and Environment
On April 20, 2010, on the Deepwater Horizon drilling rig killed 11 people and
gushed 560–585 thousand tons of crude oil into the Gulf of Mexico. On July 2, 2015,
a settlement was announced between BP, United States, and five US states. BP agreed
to pay up to US$18.7 billion in penalties [64].
18.8.13.1 Context
Deepwater Horizon floated in 4992 feet of water just beyond the continental shelf
in the Mississippi Canyon. Below the seabed was a large reservoir of oil and gas. A
long well—it was 2.5 miles long—had been drilled through the subsurface rock to
the reservoir.
The well had just been cemented. After a well is drilled, it is completed by casing
the well bore with steel pipe and cementing the casing into place. Casing prevents the
well from collapsing. Cementing fills the otherwise empty annular space between
well bore and the casing. If the annular space is totally blocked, oil and gas can flow
out of the formation only through the casing. At the top of the casing, valves, blowout
preventers, etc., control the flow of oil and provide a route for future activities, such
as perforation and well stimulation.
Figure 18.11 illustrates how centralizers improve cementing. They keep the casing
from resting against the surrounding rock, possibly blocking the flow of cement and
jeopardizing the well-completion process. According to industry acquaintances, poor
cement jobs are the leading cause of post-completion blowouts. Among other things,
poor cementing can result from using low-quality cement and not using enough
centralizers. A Halliburton model called for 21 centralizers. But there were only 6 on
the rig. In short order, 15 more were flown to the rig, but they were conventional rather
than custom-made. A BP engineer informed management that the new centralizers
were conventional, and that it would take 10 h to install them. In the end, BP installed
only six centralizers.
To test the integrity of cementing, two important tests are conducted: a positive-
pressure test and a negative-pressure test. During the positive-pressure test, the pres-
sure is increased in the steel casing and seal assembly to see if they are intact. The
negative-pressure test determines the integrity of the cement at the bottom of the
hole. Pressure is reduced to 0 psi and well evacuation is stopped. If pressure remains
low, the test is deemed successful. If flow occurs or the pressure increases, reservoir
fluids are entering the well, probably through gaps in the casing and/or the cement.
A failed negative pressure test requires remedial work to re-establish well integrity.
If the remedial work is not done, flow through well will be impossible to control.
18.8 Significant Accidents and Near-Misses 443
Rock
Centralizer
At about noon on the day of the blow-out, a helicopter arrived. Among its passengers
were four VIPs from Houston—two from Transocean, the owner of the drilling rig,
and two from BP. The executives were there to conduct a 24-hour management
visibility tour. Their presence contravened two lessons that should have been learned
from the BP Texas City disaster. During a highly critical activity, (1) non-essential
personnel must be removed from the vicinity, and (2) supervisory oversight must
be provided by at all times by technically qualified personnel. Few activities are
more critical than well completion, and few people are less essential during well
completion than corporate executives. These particular VIPS were planning to stay
on the rig overnight. They had to be fed, watered, and entertained. Their presence
contributed to the flouting of Item (2) by monopolizing the time of highly capable
experts, who otherwise might have been paying more attention to the drilling crew.
At 5:00 PM, the rig crew began the negative-pressure test. The pressure repeatedly
built back up after gases were evacuated. Between 6:00 PM and 7:00 PM, the crew
discussed how to proceed. A senior BP engineer insisted on another negative pres-
sure test. An assistant tool pusher said pressure could be measured through another
line—the “kill line”—where the pressure did indeed stay low. For whatever reason,
444 18 Safety and Environment
the crew decided that no flow through the kill line equalled a successful negative
pressure test.
The crew proceeded, preparing to set a cement plug deep in the well. They began
to replace dense drill mud, which had counteracted reservoir pressure during drilling,
with sea water. Sometime later, people on the bridge of the rig felt a high-frequency
vibration and heard a hissing noise. The senior BP engineer got a call from the
assistant tool pusher, who told him that mud was coming out of the well and that
gas was coming up too. The senior BP engineer grabbed his hard hat and stepped
outside, where saw that mud and seawater were “blowing everywhere.” The assistant
tool pusher called the senior tool pusher and said: “We have a situation … the well
is blown out.”
Soon thereafter, at roughly 9:45 PM, an explosion shook the rig, starting a fire.
A second explosion followed, then several more. Meanwhile, people were being
evacuated into life rafts. Others simply jumped into the water.
Eleven people died. Over the next several months, the blown-out well gushed
560–585 thousand tons of oil into the Gulf of Mexico, causing economic and eco-
logical harm throughout the region, from Mexico to Florida.
At about 11:00 PM on July 5, 2013, a 74-car freight train carrying Bakken crude oil
stopped in Nantes, Quebec. Nantes is about 7 miles west of Lac-Mégantic. The train
was owned by the MM&A, the Montreal, Maine and Atlantic Railway, for which
Nantes is a crew-change point.
The train was named MMA-2. With a length of 1433 m (4700 feet), it stretched
for almost a mile. Pulled by five locomotives, it weighed more than 11,000 short
tons. Each of the 72 tank cars carried about 30,000 gallons (714 barrels) of crude
oil. The oil was produced in North Dakota and was being shipped to the Irving Oil
Refinery in Saint John, New Brunswick.
While riding in his usual taxi from the train to his usual hotel, the engineer
expressed concern about the lead locomotive, because it was smoking and leaking oil.
Per MM&A policy, he had left the engine running, mainly to power the compressor
that maintained pressure in the air brakes.
At about 11:45 pm, the troubled engine caught fire. Local residents called 911,
and within five minutes, firefighters and police arrived. They saw dark smoke coming
from the engine, where a broken piston had started a good sized blaze.
After dousing the fire, a Nantes Fire Department member contacted the rail traffic
controller for MM&A in Farnham, Quebec, 125 miles away. The controller sent two
maintenance workers to the scene. They arrived at 12:13 AM on July 6. It wasn’t
safe to restart the damaged engine, so they decided to leave it off and set hand brakes
to augment the air brakes.
Without the compressor, the air brakes slowly released. The hand brakes alone
couldn’t restrain the 11,000-ton train. It began to roll toward Lac-Mégantic down a
1.2% grade. When it arrived at about 1:15 AM, the train was traveling at high speed.
It derailed. The oil spilled out. The consequent fire and explosion killed 47 people.
The glut of oil being produced from the Bakken Shale in North Dakota is putting
tremendous stress on truck and rail transportation. Old and poorly maintained equip-
ment is being pressed into service. Until the equipment is replaced with new equip-
ment or a pipeline—or oil production decreases—the stage is well-set for future
accidents.
446 18 Safety and Environment
On Wednesday, August 12, 2015, at around 22:50 local time, flames were reported in
a warehouse in the Binhai district of Tianjin, China. Situated on Bohai Bay 150 km
southeast of Central Beijing, the Binhai New Area is a Special Economic Zone
(SEZ), similar to those in Shenzhen and Pudong in Shanghai. Several international
companies, including Motorola and Airbus, have constructed facilities in the Binhai
SEZ; in 2009, Airbus opened an assembly plant there for A320 airliners.
The warehouse was owned by Ruihai Logistics. It occupied 46,000 m2 (500,000
square feet) and included multiple areas for storing containers of hazardous chem-
icals. A regulation specified that such warehouses must be located at least 1 km
away from public facilities, but in this case the regulation was ignored. Local offi-
cials, including emergency responders, did not know for sure what was stored at the
site. The situation was aggravated by major errors, including errors of omission, in
customs documents.
According to post-mortem reports, the warehouse held more than 40 kinds of
hazardous chemicals, including toluene diisocyanate, “vast quantities” of calcium
carbide (CaC2 ), 800 tons of ammonium nitrate, 700 tons of sodium cyanide, and
500 tons of potassium nitrate. CaC2 reacts with water to form calcium hydroxide
and acetylene, a highly flammable gas. Sodium cyanide reacts with water to form
sodium hydroxide and hydrogen cyanide gas, which is both toxic and flammable.
Ammonium nitrate is the chemical responsible for the massive explosion in Texas
City, Texas in 1947.
First responders to the fire were unable to keep it from spreading. Unaware of
what they were facing, they tried dousing the flames with water. The consequent
explosion, possibly due to acetylene, was equivalent to 3 tons of TNT. Less than a
minute later, the ammonium nitrate detonated, causing a blast equivalent to 20 tons of
TNT. The blaze continued for several days, in part because of a pause in firefighting
due to lingering uncertainty about the kinds and quantities of chemicals at the site. On
August 15, the spreading flames triggered eight more explosions, which, fortunately,
were relatively minor.
Photographs showed extensive destruction around the warehouse, with a massive
crater at the center. Seven nearby buildings were destroyed, and others were rendered
structurally unsafe. More than 17,000 nearby apartment units were damaged; 6000
people were evacuated.
On August 18, the first rain after the fire coated nearby streets with white chemical
foam. People complained that contact with the rain caused burning sensations and
rashes.
Also on August 18, the Central commission for Discipline Investigation (CCDI)
initiated an investigation of Yang Dongliang. Yang was Director of the State Admin-
istration of Work Safety—China’s highest-ranking work safety official. He had been
Tianjin’s vice mayor for 11 years. In 2012, he had issued an order that loosened rules
18.8 Significant Accidents and Near-Misses 447
for the handling of hazardous substances. On October 16, he was expelled from the
Communist Party for corruption and other offences.
On August 20, tons of dead stickleback fish washed up on the land about 6 km from
the explosion site. Officials downplayed the danger, claiming that the fish probably
died due to oxygen depletion, not cyanide poisoning.
On August 25, another rain brought more complaints of skin burns, and white foam
again appeared on the streets. The director of Tianjin’s environmental monitoring
center claimed that the foam was “a normal phenomenon when rain falls. Similar
things have occurred before.”
According to a report issued on September 12, the human toll at Tianjin was 173
killed, 797 hospitalized, and 8 still missing.
During the past several years, in my technical support role, I (PRR) have witnessed
shutdowns, near-misses, and loss of containment due to known (but un-amended)
process design flaws, inadequate maintenance, inadequate training, lack of technical
supervision, contravention of procedures, and worker fatigue. DIBS—Did It Before
Syndrome—played a role in three of the four examples in this section, and in many
of the examples described above. DIBS leads workers and managers to disregard
safety rules and procedures, because they got away with doing so before.
The refinery in which this incident occurred was owned by a company known for its
dedication to safety. Behind the hydrocracker area in the control room was a large
sign that said, in effect, “All employees are obliged to stop any operation he or she
believes is unsafe.” At midnight, it was time for the initial wetting of the catalysts
with oil.
Initial wetting occurs once and lasts about an hour, but it releases heat due to
enthalpy of adsorption. In multiple-bed hydrocrackers (see Fig. 18.9) wetting causes
the Bed 1 outlet temperature to rise, within minutes, by as much as 50 °C. Wetting
exotherms can be just as high in subsequent beds. Oil adsorbs heat, so the oil rate
is kept high, usually between 50 and 70% of the design maximum. If the oil rate is
lower, the temperature rise is higher.
Operators can prepare for the exotherms with post-quench. When oil reaches the
inlet to Bed 1, they start sending cold gas into the quench deck between Bed 1 and
Bed 2. It can take 5–8 min for oil to pass through the bed. When the oil arrives at
the Bed 1 outlet, it mixes with the cold post-quench gas, which quickly drops the
temperature. Without post-quench—that is, if operators didn’t add cold gas until
they saw the exotherm—the temperature wave would pass through the quench box
448 18 Safety and Environment
Fig. 18.12 Comparison of oil temperatures during initial catalyst wetting, with and without post-
quench. The black and gray segments are for temperatures in the quench boxes. The “no Q” trends
do not include post-quench
to the inlet of Bed 2 unabated. The heat of adsorption in Bed 2 would increase the
temperature by another 50 °C. The total increase across in Beds 1 and 2 could be
100 °C. And so on.
Figure 18.12 compares wetting with and without post-quench. Without post-
quench, wetting heat accumulates. Industry experience shows that temperatures can
get high enough to initiate cracking and even temperature excursions. After the cat-
alyst is wet, temperatures quickly fall to pre-wetting level.
Just after midnight in the subject refinery, the hydroprocessing Area Manager,
unit engineer, were discussing how to proceed despite some known problems. As
at BP Texas City, there were three known problems in a separation tower: the level
transmitter wasn’t working, the sight glass was foggy, and the offgas compressor
wasn’t working. There was no way to observe level or control pressure in the tower.
Repairs work couldn’t start before 8:00 AM, and then it would take an unknown
time to fix the compressor. An in-house technician probably could repair the level
controller in parallel with other work.
Two VIPs—the refinery Engineering Manager and Operations Superintenden-
t—appeared in the control room and immediately started firing questions. They were
concerned that wetting was being delayed. They asked operator to show panel dis-
plays irrelevant to the job at hand. Observing the operator’s distress, I directed the
VIPs to a nearby conference room. The room contained an observation-only panel
18.9 Personal Observations 449
display, with which they could browse without disturbing the startup team. When I
returned to the control room, the unit operator and operator thanked me effusively.
The Area Manager decided to proceed despite the known problems. He concluded
that with post-quench operation, the catalyst wouldn’t get hot enough to induce
hydrocracking. When wetting was over, oil could be circulated for a few hours to
flush particulates out of the catalyst. By then, it would be 7:00 AM and the unit could
be brought down for repairs.
Wetting started at about 3:00 AM. When liquid feed was introduced, a 4th problem
became apparent: the feed pump valve was “sticky,” taking up to three minutes to
respond to a change in setpoint. Inability to control feed was crucial, due to the impact
of feed on the size of exotherms. The operator called across the control room to the
Area Manager. “This is not safe,” he said. “I want to shut it down.” According to
company policy, he should have brought the unit down immediately, without asking
permission. The
Area Manager said: “Steady as she goes. We can’t go back. It’s okay. We’ve done
this before.”
During wetting, the oil flow was lower than specified. With less oil to pick up
heat, exotherms were larger than ever before. But thanks to post-quench operation,
there was never any danger of a temperature excursion.
However, the downstream tower overflowed, sending hydrocarbons to the flare.
The flaring lasted an hour or so, but it had to be reported to environmental authorities.
The startup was delayed for days. Saving five hours cost the refinery more than $2
million.
The Area Manager had gotten away with operating the tower before without
the overhead compressor. He had also gotten away with running without the level
controller. He had never run the unit with both of those problems while also lacking
reliable control of feed rate.
As for me (PRR): I was consulted by the operator during the event, but not by
his boss. I did report the event to friend, a corporate subject matter expert. He was
appalled. I never heard what action he took.
A serious excursion was the latest of several in a new hydrocracker. The ultimate
cause was a decision by the purchasing department to order substandard items for
the associated hydrogen plant. One by one, the faulty items failed and were replaced
with proper items. Meanwhile, during its first 6 months of operation, the unit never
ran continuously for more than one month.
Each hydrogen plant failure upset the hydrocracker. Some failures sent liquid
into the suction of the hydrocracker recycle gas compressor (RGC), causing the
RGC to shut down. The licensor’s EDS interlock gave operators 15 min to restart
450 18 Safety and Environment
Fig. 18.13 Pressure, recycle gas flow, and selected temperatures for Example 2. Bed Outlet TI-
1 flat-lines at 618 °C, because 618 °C is top-of-range. Actual temperatures were far higher, as
evidenced by the fused catalyst in the bed when the reactor was unloaded
the compressor before automatically activating the low-rate depressuring valve. Each
depressuring event shut the unit down for several hours if not days. To give themselves
extra time to restart the RGC, the unit engineer and operators decided to disable
the EDS/RGC interlock, assuming that depressuring triggered by high temperature
would be sufficient.
This demonstrated ignorance of several fundamental concepts. The most egre-
gious were:
• Residence time has a significant impact on reaction rates.
• The flow of fluids is the only significant way to remove heat from a hydroprocessing
reactor.
• With zero or low flow, bed outlet TI readings are meaningless.
It is crucial to understand the difference between TI readings and temperature. A
previous example shows that one never knows the highest or lowest temperature in a
catalyst bed. But when reactants are flowing through at 50–100% of design capacity,
one at least knows that bed outlet TIs correspond to the temperatures above them.
When flow stops, no TI reading is relevant to any temperature more than a few inches
away. Reactions do not stop when flow stops. Heat accumulates in catalyst beds, often
without affecting the bed outlet TI readings. Without gas flow, hot oil pools due to
gravity at low points. Eventually, the pooling oil will reach one or more bed outlet
TIs and/or reactor wall TIs, providing the first tangible indication that something is
wrong. Undoubtedly, that is what happened in this unit. Table 18.6 shows a timeline
for this incident. A visual representation is presented in Fig. 18.13.
18.9 Personal Observations 451
After the first compressor failure, it took 144 min for some overheated oil to
affect TI reading at the reactor outlet. At 17:38 h, an upward spike in Reactor Outlet
TI-3 triggered depressuring when it reached 440 °C. TI-3 continued to rise until it
reached 487 °C. Depressuring seemed to make things worse. It decreased TI-3, but as
the pressure fell, the exiting gas pulled formerly stagnant hot oil through the reactor.
Between 17:48 and 17:55 h, Bed Outlet TI-1 jumped from 404 to 618 °C, an increase
of 214 °C in 7 min.
The readings dropped after a restart of the RGC. But then, most likely due to the
arrival of more overheated oil, TI-1 jumped back up to 618 °C, where it flat-lined
for more than 2½ hours. Actual temperatures were higher than 618 °C, which is
top-of range. It was thought that the TI-1reading was faulty, but it wasn’t; it returned
to reading on-scale at 20:54. A second attempt to restart the RGC succeeded. TI-1
readings started to fall. But as it did, Bed Outlet TI-2 rose above 550 °C and stayed
there for another two hours.
Afterwards, the unit ran at reduced rates for six months, with strict adherence
to licensor-suggested temperature limits. The limits were tighter due to fears that
the reactor outlet pipe had been compromised. During the subsequent turnaround,
452 18 Safety and Environment
workers found that the reactor internals were damaged. They also found partially
fused silica balls near the reactor wall. Pure silica melts at about 2000 °C. Impurities
drop the fusion temperature considerably. But even so, it must have gotten incredibly
hot in that reactor.
Refineries schedule maintenance shutdowns for several units at once, often more than
a year in advance. Each process unit depends on several others, serving as a destina-
tion for some streams or as a source of feedstock for other units. When a hydrocracker
comes down, the need for hydrogen goes down, and major units normally fed by the
hydrocracker have to be run differently. To do special work, equipment replacements
must be ready to install, catalysts must be onsite, oil inventories must be adjusted, and
maintenance contractors are hired to conduct repairs, provide maintenance, unload
and load catalysts, etc. An unplanned shutdown is far more expensive than a planned
shutdown, because other units aren’t prepared, and maintenance personnel may not
be instantly available.
The following instance was caused (mainly) by a decision to reduce the number
of crews employed to service heat exchangers. In shell-and-tube exchangers, one gas
or liquid stream is pumped through a bundle of parallel tubes. As it flows through
a bundle, the “tube-side” fluid exchanges heat with “shell-side” fluid, which flows
in the opposite direction, inside the shell but outside the tubes. A key activity is the
cleaning or replacement of fouled tubes. Due to the shortage of people, exchanger
maintenance was taking a lot longer than expected. It was clear that the pre-ordained
schedule could not be met, given the original scope of work. A decision was made to
clean just the exchangers most in need of cleaning. For an important set of exchangers
at the hydrocracker, the feed/effluent exchanger (F/E), the crew cleaned only some
tubes—the ones most badly fouled. Upon restart, flow was even in the F/E, where
it preferentially followed the path of least resistance—most likely through the clean
tubes. Within weeks, under-performance of the F/E required an unplanned hydro-
cracker shutdown and costly changes in the operation other units. The shutdown,
service work, and subsequent restart required about two weeks at a cost of more than
US$ 1 million per day—a steep price to pay for saving less than $50,000 per day
from reducing tube-cleaning personnel.
The decision to clean only some tubes in the F/E displayed fundamental ignorance.
Not knowing about the crew’s decision revealed slip-shod management. No wonder.
The refinery was chronically understaffed. During maintenance shutdowns, engineers
had to work as many as 30 consecutive 14-hour days. Staffing decisions are made by
refinery management, who are ultimately responsible for the consequences.
18.9 Personal Observations 453
Despite the harsh comments in the BP Texas City report about worker fatigue and
insufficient training, a small but significant number of U.S. refineries remain under-
staffed, especially for shutdown and startup activities. Those refineries aren’t alone.
Many U.S. service companies also are under-staffed, especially during “turnaround
season” in March through May. Refineries want to complete maintenance before
June, so they can be up and running throughout the summer, when gasoline con-
sumption is highest. Members of a vendor technical service team had been working
long hours during the spring startup season, often more than 100 h per week. In addi-
tion to spending 12–14 h per day supervising startups, team members were required
to handle questions from other customers, write proposals, and prepare for upcoming
customer visits. Four hours of sleep per night was the rule, not the exception.
The group had always been busy, but the workload increased after the business
unit was reorganized. The main reason was the switch from a leader who managed
the team’s workload by screening sales-support work, to a person who chased every
apparent sales opportunity, no matter how unlikely.
One night, while driving from a refinery to his hotel, a team member fell asleep
and drove off a country road. He wasn’t injured. For that reason and others, he didn’t
report the incident until several months later, when another member of the team
persuaded him to complain.
In response to the complaint, which included mention of possible consequences, a
high-level manager said: “What are you, a lawyer? Mister HSE?” followed by “Why
didn’t you say something sooner?” The manager later admitted knowing about the
over-work for several months and had considering adding staff. “We didn’t,” the
high-level manager said, “because we might have had to lay people off if business
slowed down.” Within a year of the reorganization, 60% of the original team had
left, primarily due to over-work.
The high-level manager’s response to what could have been a fatal accident was
completely inconsistent with company policy. Accidents and near-miss reports were
supposed to be treated seriously, no matter how or when they were reported. The
fact that the manager knew about the over-work for so long—and did nothing—is
astounding.
18.10 Lessons
Lessons are learned such events. But not everyone learns. Many of those who do
learn forget. Or worse, they decide that the lessons don’t apply to them.
454 18 Safety and Environment
18.10.1 Dibs
injury, 80,000 trips without dying, and 270,000 trips without dying due to speeding.
So we speed, even though we know speeding is dangerous.
In hydrocarbon industries, accidents can be far worse, threatening the health and
prosperity of thousands of people. In well-run facilities, the personal financial cost
of breaking a rule can far exceed the cost of a speeding ticket; it can get a worker
fired.
But as described in several of the examples above, workers, managers and exec-
utives—knowingly and willfully—continue to break critical rules. Why? One can
only guess. Is it simply because the rules are inconvenient? Do they believe the rules
don’t apply to them? Or is it DIBS—concluding that their way is better, because
they’ve done it that way before with no negative consequences?
How can DIBS be diminished? Experience shows that the most effective way to
decrease DIBS is to threaten individuals. If people can lose a job or go to jail for
violating a rule, they take the rule more seriously, whether or not they agree with it.
18.10.2 MOC
Perhaps a DIBS practitioner is correct. Maybe his/her way really is okay, or even
better.
Management of Change (MOC) [71] is a formal process by which procedures can
be improved. With design models, a written document, and perhaps a formal pre-
sentation, the originator presents an idea to everyone involved. For minor, localized
change in procedure, the MOC process can be quick, requiring approval from two or
more colleagues and subsequent documentation. For a significant change in hardware
or general operating procedures, an MOC committee usually includes peers, super-
visors, and internal experts. It might also include one or more outside consultants.
Implementation might require detailed design, budget approval, and development
and review of operating procedures.
Ineffective MOC is one of the leading causes of serious incidents, according to the
U.S. Chemical Safety and Hazard Investigation Board (CSB) [72]. CSB stated: “In
industry, as elsewhere, change often brings progress. But it can also increase risks
that, if not properly managed, create conditions that may lead to injuries, property
damage or even death.”
Changes subject to MOC procedures can include:
• Personnel changes, including reorganization, staffing, and training.
• Equipment changes, including revamps, maintenance schedules, communications
devices, control hardware, and control software.
• Procedure changes, including procedures for normal operation, startup, shutdown,
and MOC.
• Material changes, including significant changes in catalysts, chemicals, and feed-
stock.
456 18 Safety and Environment
nobody in the plant understood the relationship between N-slip and hydrocracking
catalyst activity.
At BP Texas City, operators proceeded with the startup despite the lack of critical
level measurement, reflecting a combination of inadequate training, inadequate tech-
nical supervision, inadequate communications between shifts, and probably DIBS.
Similar factors played a role in the Deepwater Horizon incident. Fatal mistakes
included the use of an insufficient number of centralizers and the assumption that
pressure in the kill line was indicative of pressure in the well. The drilling crew did
not understand the significance several consecutive failed negative pressure tests.
The visit of non-essential VIPs during the most critical phase of well completion
distracted key technical personnel, who quickly would have known the significance
of the negative-pressure test.
The train derailment in Lac-Mégantic was caused by faulty maintenance, the
failure of the engineer to tell MM&A about the failing locomotive before it caught
fire, and the failure of authorities to act to ensure that oil transportation infrastructure
is safe.
All of the other examples occurred during startups. All were influenced by cost-
cutting and/or unreasonable schedules. The most inexcusable are those that were
caused by rogue managers who countermanded company policy in companies that
spend millions on health and safety training.
Of these cases:
• Seven occurred during a startup or shutdown.
• Nine were caused in large part by contravention of procedures. Of these, three
involved willful disabling of safety interlocks.
• Eight were caused by cost-cutting.
• Eight were aggravated by inadequate training.
This section focuses almost exclusively on U.S. environment protection agencies and
laws. Please understand that to include similar detail for other countries would have
been infeasible in the space allotted.
In 1970, United States formed the Environmental Protection Agency (US EPA)
[76] and the Occupational Safety and Health Administration (OSHA). Together,
these agencies are responsible for dramatic improvements in air and water quality
and increases workplace safety throughout the United States.
EPA is an amalgamation of departments, bureaus, agencies, etc., taken from the
Interior Department; the Department of Health, Education, and Welfare; the Agricul-
ture Department; the Atomic Energy Commission; the Federal Radiation Council;
and the Council on Environmental Quality. EPA’s mission is “to enforce federal laws
to control and abate pollution of air and water, solid waste, noise, radiation, and toxic
substances. It is also to administer the Superfund for cleaning up abandoned waste
sites, and award grants for local sewage treatment plants.”
After its creation, EPA quickly took the following actions:
• Established 10 regional offices throughout the nation
• Established National Ambient Air Quality Standards, which specified maximum
permissible levels for major pollutants
• Required each state to develop plans to meet air quality standards
• Established and enforced emission standards for hazardous pollutants such as
asbestos, beryllium, cadmium, and mercury
• Required a 90% reduction in emissions of VOC and carbon monoxide by 1975
• Published emission standards for aircraft
• Funded research and demonstration plants
• Furnished grants to states, cities, and towns to help them combat air and water
pollution.
EPA’s law-enforcement efforts are supported by the National Enforcement Inves-
tigation Center in Denver, Colorado, which gives assistance to federal, state, and local
law enforcement agencies. This unit has clamped down on the “midnight dumping”
of toxic waste and the deliberate destruction or falsification of documents.
Today, almost every U.S. state has an environmental agency. Arguably, the most
famous of these is the California Air Resources Board (CARB), which pioneered
regulations to mitigate smog in Los Angeles. In addition to administering state pro-
grams for improving air and water quality, the Texas Commission on Environmental
Quality (TCEQ) participates in making plans to prevent and react to industrial ter-
rorism.
460 18 Safety and Environment
Most countries have environmental agencies. As shown in Table 18.7, some are
combined with public health departments, some are combined with energy agencies,
and at least one is coupled with tourism. In addition to handling internal issues, most
of these agencies administer their country’s participation in international treaties,
such as the Kyoto Protocol [77].
18.13 Occupational Safety and Health Administration 461
Pollution control and safety are two sides of the same coin. In the United States, the
Occupational Safety and Health Administration (OSHA) is part of the U.S. Depart-
ment of Labor. Its legislative mandate is to assure safe and healthful working condi-
tions by:
• Enforcing the Occupational Safety and Health Act of 1970
• Helping states to assure safe and healthful working conditions
• Supporting research, information, education, and training in occupational safety
and health.
OSHA can levy fines against unsafe people and companies. Not surprisingly, a
large percentage of the safety infringements cited by OSHA have caused environ-
mental damage or put the environment at risk.
• In 1993 OSHA fined the Manganas Painting Company for exposing workers
and the environment to lead during sand-blasting operations. The proposed fines
totalled US$4 million. The contractor appealed, but in February 2002, it pled
guilty to knowingly and illegally dumping 55 tons of lead-containing sandblasting
material, in violation of the Resource Conservation and Recovery Act (RCRA)
[78].
• BP was fined a then-record US$21 million for infractions related to the 2005 Texas
City explosion. After a 2009 followup, OSHA found that BP had failed to abate a
number of safety-related items and committed several willful violations of safety
regulations. In 2010, BP agreed to pay US$50.6 million to settle the failure-to-
abate violations and in 2012 it agreed to settle 409 of 439 willful violations for
US$13 million.
BP pled guilty to 11 felony counts of misconduct or neglect and paid US$4.5 bil-
lion for the Deepwater Horizon incident. The payment included more than US$1.25
billion to settle the criminal fines.
Under the Occupational Safety and Health Act, employers are responsible for ensur-
ing a safe and healthy workplace. One key requirement is that all hazardous chemicals
must be properly labeled. Workers must be taught how to handle the chemicals safely,
and material safety data sheets must be available to any employee who wishes to see
them.
Material Safety Data Sheets (MSDS) include the following information:
Material identification. The name of the product and the manufacturer’s name,
address, and emergency phone number must be provided.
462 18 Safety and Environment
Hazardous ingredients. The sheet must give the chemical name for all hazardous
ingredients comprising more than 1% of the material. It must list cancer-causing
materials if they comprise more that 0.1%. Listing only the trade name, only the
Chemical Abstract Service (CAS) number, or only the generic name is not acceptable.
If applicable, exposure limits are listed in this section of the MSDS. The OSHA
permissible exposure limit (PEL) is a legal, regulated standard. Other limits may also
be listed. These include recommended exposure limits (REL) from the National Insti-
tute for Occupational Safety and Health (NIOSH) and threshold limit values (TLV)
from the American Conference of Governmental Industrial Hygenists (ACGIH).
Sometimes, short-term exposure and/or ceiling limits are shown. The ceiling limit
should never be exceeded.
Physical properties. These include the appearance, color, odor, melting point,
boiling point, viscosity, vapor pressure, vapor density, and evaporation rate. The
vapor pressure indicates whether or not the chemical will vaporize when spilled. The
vapor density indicates whether the vapor will rise or fall. Odor is important because
a peculiar smell is the first indication that something has leaked.
Fire and explosion hazard data. This section provides the flash point of the mate-
rial, the type of extinguisher that should be used if it catches fire, and any special
precautions.
The flash point is the lowest temperature at which a liquid gives off enough vapor to
form an explosive mixture with air. Liquids with flash points below 100 °F (37.8 °C)
are called flammable, and liquids with flash points between 100 and 200 °F (37.8 and
93.3 °C) are called combustible. Flammable and combustible liquids require special
handling and storage.
The four major types of fire extinguishers are Class A for paper and wood, Class
B for flammable liquids or greases, Class C for electrical fires, and Class D for fires
involving metals or metal alloys.
Health hazard data. This section defines the symptoms that result from normal
exposure or overexposure to the material or one of its components. Toxicity informa-
tion, such as the result of studies on animals, may also be provided. The information
may also distinguish between the effects of acute (short term) and chronic (long-term)
exposure. Emergency and first-aid procedures are included in this section.
Reactivity data. This section includes information on the stability of the material
and special storage requirements.
Unstable chemicals can decompose spontaneously at certain temperatures and
pressures. Some unstable chemicals decompose when they are shocked. Rapid
decomposition produces heat, which may cause a fire and/or explosion. It also may
generate toxic gas. Hazardous polymerization, which is the opposite of hazardous
decomposition, also can produce enough heat to cause a fire or explosion.
Concentrated acids and reactive metals are hazards when mixed with water. They
should be stored separately in special containers.
Spill, leak, and disposal procedures. This part of the MSDS gives general proce-
dures, precautions and methods for cleaning up spills and disposing of the chemical.
18.13 Occupational Safety and Health Administration 463
The first clear-air acts in the English-speaking world were implemented by Parliament
in 1956, in response to the “deadly fog” incidents around London.
Since 1963, the United States government has passed several clean-air acts, includ-
ing:
• Clean Air Act of 1963
• Motor Vehicle Air Pollution Control Act of October 20, 1965
• Clean Air Act Amendments of October 15, 1967
• Air Quality Act of November 21, 1967
• Creation of EPA: Clean Air Act of 1970
• Clean Air Act Amendments of 1975, 1977, and 1990.
464 18 Safety and Environment
For convenience, the entire package often is called just the Clean Air Act (CAA).
A good source for general understanding is the Plain English Guide to the Clean Air
Act, available from the U.S. EPA web site [80]. The Guide shows how the CAA has
led to significant improvements in human health and the environment. For example,
since 1970:
• The six most commonly found air pollutants have decreased by more than 50%
• Air toxics from large industrial sources, such as chemical plants, petroleum refiner-
ies, and paper mills have been reduced by nearly 70%
• New cars are more than 90% cleaner and will be even cleaner in the future, and
• Production of most ozone-depleting chemicals has ceased
At the same time, the U.S. economy and associated energy consumption has grown
significantly:
• The U.S. GDP has tripled
• Energy consumption has increased by 50%
• Vehicle use has increased by almost 200%.
Under the CAA, EPA’s role is to set limits on certain air pollutants, including
ambient levels anywhere in the country. EPA limits emissions of air pollutants from
point sources like chemical plants, utilities, and steel mills. Individual states or tribes
may have stronger air pollution laws, but they may not have weaker pollution limits
than those set by EPA.
EPA assists state, tribal, and local agencies by providing research, expert studies,
engineering designs, and funding to support clean air progress. EPA must approve
state, tribal, and local air pollution reduction plans. EPA can issue sanctions against
states which fail to comply; if necessary, it can take over enforcing the CAA in
non-complying areas.
The CAA required EPA to set National Ambient Air Quality Standards (NAAQS)
for pollutants considered harmful to public health and the environment. EPA set
two types of standards—primary and secondary. Primary standards protect against
adverse health effects. Secondary standards protect against damage to farm crops,
vegetation, and buildings. Because different pollutants have different effects, the
NAAQS are also different. Some pollutants have standards for both long-term and
short-term averaging times. The short-term standards are designed to protect against
acute, or short-term, health effects, while the long-term standards were established
to protect against chronic health effects. NAAQS are shown in Table 18.8 [81].
Hazardous air pollutants (HAPs) are regulated, too. Table 18.9 lists 20 “core”
compounds selected from of the 188 HAPs regulated under Sect. 112 of the Clean
Air Act.
Compliance with air quality standards is monitored by the Office of Air Quality
Planning and Standards (OAQPS) [81]. OAQPS evaluates the status of the atmosphere
as compared to clean air standards and historical information, using measurements
acquired from many hundreds of monitoring stations across the United States.
18.14 Key Regulations 465
In 1970, gasoline blending became more complex. The U.S. Clean Air Act required
the phase-out of tetraethyl lead, so refiners had to find other ways to provide octane.
In 1990, the CAA was amended. It empowered EPA to impose emissions limits on
automobiles and to require reformulated gasoline (RFG). RFG was implemented in
several phases. The Phase I program started in 1995 and mandated RFG for 10 large
metropolitan areas. Several other cities and four entire States joined the program
voluntarily. The current phase began in 2000. RFG is used in 17 States and the
District of Columbia. In 2015, about 35% of the gasoline in the United States was
reformulated.
Tier 1 reformulated gasoline regulations required a minimum amount of chem-
ically bound oxygen, imposed upper limits on benzene and Reid Vapor Pressure
(RVP), and ordered a 15% reduction in volatile organic compounds (VOC) and air
toxics. VOC react with atmospheric NOx to produce ground-level ozone. Air toxics
include 1,3-butadiene, acetaldehyde, benzene, and formaldehyde.
The regulations for Tier 2, which took force in January 2000, were based on
the EPA Complex Model, which estimates exhaust emissions for a region based on
geography, time of year, mix of vehicle types, and—most important to refiners—fuel
properties. As of 2006, the limit on sulfur in the gasoline produced by most refineries
in the U.S. was 30 ppmw.
Initially, the oxygen for RFG could be supplied as ethanol or C5 –C7 ethers. The
ethers have excellent blending octanes and low vapor pressures. But due to leaks from
filling station storage tanks, methyl-t-butyl ether (MTBE) was detected in ground
water samples in New York City, Lake Tahoe, and Santa Monica, California. In
1999, the Governor of California issued an executive order requiring the phase-out
of MTBE as a gasoline component. That same year, the California Air Resources
Board (CARB) adopted California Phase 3 RFG standards, which took effect in stages
18.14 Key Regulations 467
starting in 2002. The standards include a ban on MTBE and a tighter cap on sulfur
content—15 ppmw maximum. In the United States, Tier 3 vehicle emission and fuel
standards lowered the allowed sulfur content of gasoline to 10 ppmw, beginning in
2017.
The Energy Policy Act of 2005 [83] was signed into law by President George W.
Bush on August 8, 2005. The Act addresses energy production and use in the United
States. It provides tax incentives and loan guarantees for specific types of energy
production Topics include:
• Energy efficiency
• Renewable energy
• Oil and gas
• Coal
• Tribal energy
• Nuclear matters and security
• Vehicles and motor fuels, including ethanol
• Hydrogen
• Electricity
• Energy tax incentives
• Hydropower and geothermal energy
• Climate change technology.
Critics say that the Act is a broad collection of subsidies for United States energy
companies; in particular, the nuclear and oil industries [84]. It authorized the follow-
ing tax reductions:
• $4.3 billion for nuclear power
• $2.8 billion for fossil fuel production
• $2.7 billion to extend the renewable electricity production credit
• $1.6 billion in tax incentives for investments in “clean coal” facilities
• $1.3 billion for energy conservation and efficiency
• $1.3 billion for alternative fuels (bioethanol, biomethane, liquified natural gas,
propane)
• $500 million Clean Renewable Energy Bonds for government agencies for renew-
able energy projects.
The bill exempted hydraulic fracturing from certain provisions of the Safe Drink-
ing Water Act. Consequently, drilling companies don’t have to disclose the chemicals
used in fracking.
468 18 Safety and Environment
The most questionable provision of the Act increases the amount of biofuel
(specifically corn ethanol) that must be blended with gasoline. Proponents say that
displacing gasoline with ethanol achieves significant CO2 production and decreases
dependence on petroleum imports. Detractors point out that while corn can give
1.1–1.4 times more energy than is required for its growth, transportation, and pro-
cessing [85], sugar cane in Brazil [86] gives an energy gain of 8.0. Large amounts
of arable land are used to grow corn for ethanol. The ethanol mandate is increasing
food prices and putting pressure on ground water. Furthermore, instead of specifying
concentrations of ethanol in the gasoline blend, the law specifies the total amount of
ethanol that must be consumed, even if gasoline sales decrease.
The Energy Independence and Security Act of 2007 required improved vehicle
fuel economy. It also increased production requirements for biofuels, specifically
biomass-derived diesel. To be biomass-based diesel, the fuel must reduce emissions
by 50% versus petroleum-derived diesel. It increased the total target volume for
ethanol, and specified that a large portion should come from non-corn sources. The
Act motivates energy savings with improved standards for appliances and lighting,
and in industrial and commercial buildings.
At present, the most advanced specification for diesel fuel is Euro V, which is
described in Table 18.10. Severe hydroprocessing is required to making Euro V
from difficult feedstocks.
One of the first environmental laws in the U.S. was the Rivers and Harbor Act of 1899.
It was passed to control obstructions to navigation. The Act required congressional
approval for the building of bridges, dams, dikes, causeways, wharfs, piers, or jetties,
either in or over a navigable waterway. The Act also made it illegal to discharge debris
470 18 Safety and Environment
into navigable water without a permit. In 1966, a court held that the River and Harbor
Act made it illegal to discharge industrial waste without a permit, not just directly
into navigable waters, but also into associated tributaries and lakes, i.e., just about
every puddle of open water in the United States. This led to the Refuse Act Permit
Program of 1970, under which specific kinds of pollution are allowed under permits
issued by the Army Corps of Engineers. Every application for a permit is reviewed
by EPA. If EPA concludes that the discharge described in the application will harm
the environment, the Army Corp of Engineers denies the permit.
The penalties for violating permits are severe, including stiff fines and jail time. A
corporation that employs the guilty person can be fined up to $1,000,000. False reports
also can be punished by fines or imprisonment. In this context, “guilty person” refers
to the corporate officer responsible for the facility from which the illegal discharge
originates. In other words, a negligent act by a sloppy operator can send the Big Boss
to jail.
The original Federal Water Pollution Control Act (FWPCA) was approved on July
9, 1956. The present Act includes the following:
• Pollution Control Act Amendments of July 20, 1961
• Water Quality Act of October 2, 1965
• Clean Water Restoration Act of November 3, 1966
• Water Quality Improvement Act of April 3, 1970
• Federal Water Pollution Control Act of 1972
• Clean Water Acts of 1977, 1981, and 1987.
The FWPCA of 1972 gave EPA greater authority to fight water pollution. While
implementing the Act, EPA cooperates with the U.S. Coast Guard and the Secretary of
the Interior. Individual states have primary responsibility for enforcing water quality
standards, but if the states fail to meet expectations, EPA can take civil or criminal
action under the Refuse Act.
The FWPCA prohibits the discharge of harmful amounts of oil into navigable
waters. If oil is spilled, the owner or operator is liable for cleanup costs. Initially, the
bill authorized $24.6 billion for water pollution control over three years. The goal of
the law was to eliminate the pollution of U.S. waterways by municipal and industrial
sources by 1985.
18.14 Key Regulations 471
The main objective of the Clean Water Acts (CWA) of 1977, 1981 and 1987 is to
maintain the “chemical, physical, and biological integrity of the nation’s waters.” It
seeks to have “water quality which provides for the protection and propagation of
fish, shellfish and wildlife and provides for recreation in and on the water.” Under
these Acts, each state is required to set its own water quality standards. All publicly
owned municipal sewage treatment facilities are required to use secondary treatment
for wastewater. To help states and cities build new or improved water treatment
plants, Congress provides construction grants, which are administered by EPA. EPA
allocates funds to states, which in turn distribute money to local communities.
Community programs are monitored by EPA. They must meet treatment require-
ments to obtain permits under the National Pollutant Discharge Elimination System
(NPDES).
The Water Quality Act of 1987 requires discharge permits for all point sources
of pollution. More than 95% of all major facilities now comply with 5-year NPDES
permits, which specify the types and amounts of pollutants that legally can be dis-
charged. When permits are renewed, they can be modified to reflect more stringent
regulations. Violators are subject to enforcement actions by EPA, including criminal
prosecution.
The authority of the EPA was strengthened under the 1987 Water Quality Act.
The allowable sizes of fines were increased, and violators found guilty of negligence
could be sent to prison.
The Marine Protection, Research, and Sanctuaries Act of 1972 gave authority to the
EPA to protect oceans from indiscriminate dumping. The Agency designates sites
at which dumping is allowed and issues dumping permits. Fines can be imposed for
illegal dumping.
Since the 1970s, the assurance of safe drinking water has been a top priority for EPA,
along with individual states and over 53,000 community water systems (CWSs). The
CWSs supply drinking water to more than 280 million Americans—about 90% of
the population.
The Safe Drinking Water Act of 1974 was amended in 1977 and again in 1986.
It empowered EPA to set national standards for drinking water from surface and
underground sources. It also authorized EPA to give financial assistance to states,
472 18 Safety and Environment
which are in charge of enforcing the standards. Aquifers are protected from wastes
disposed in deep injection wells, from runoff from hazardous waste dumps, and from
leaking underground storage tanks. In 1987, EPA also established maximum contam-
inant levels for volatile organics (VOC) and 51 manmade chemicals. Standards for
other chemicals were added as their toxicity was determined. At present, health and
safety standards have been established for 91 microbial, chemical, and radiological
contaminants.
Years ago, people were less aware of the dangers of dumping chemical wastes. On
many properties, dumping was intensive and/or continuous. This created thousands
of hazardous sites, many of which were uncontrolled and/or abandoned. On Decem-
ber 11, 1980, in response to public concern, Congress passed the Comprehensive
Environmental Response Compensation and Liability Act (CERCLA) [88], which
authorized EPA to locate, investigate, and remediate the worst of these hazardous
sites.
CERCLA established the Superfund, which provides emergency cleanup funds for
chemical spills and hazardous waste dumps. The Superfund allows the government
to take immediate action to cleanup spills or dumps where the responsible party
cannot be identified easily. The Superfund draws about 90% of its money from taxes
on oil and selected chemicals. The remainder comes from general tax revenues.
Except in an emergency, state agencies are consulted before the federal govern-
ment takes action. When it does so, it uses one of three approaches:
1. If the owner of the hazardous site cannot readily be identified, the federal gov-
ernment may proceed with the cleanup.
2. If the owner can be identified but refuses to clean the site, or if the owner’s efforts
are not up to par, the federal government can take charge of the cleanup. The
owner must pay the cost, whatever it happens to be.
3. When the owner can be identified and decides to do the work, the federal govern-
ment monitors the project and gives official approval when the work is completed
according to standards.
CERCLA covers a wide range of sites. In addition to land-based dumps, it applies
to spills into waterways, groundwater, and even the atmosphere. Initial funding for
the Act was US$1.6 billion over 5 years. In 1986, the Superfund Amendment and
Reauthorization Act (SARA) extended the program by five years and increased the
fund to US$8.5 billion. It also tightened cleanup standards and enhanced EPA’s
enforcement powers. The Emergency Planning and Community Right-to-Know Act
(EPCRA), also known as SARA Title III, encourages and supports emergency plan-
ning efforts at state and local levels. It also gives information to the public and local
governments on potential chemical hazards in their communities. This legislation
helped reduce pollution and improve safety all across the land.
In 1984, Hazardous and Solid Waste Amendments (HSWA) were passed because
citizens were concerned about the potential contamination of ground water by haz-
ardous waste disposal sites.
In 1978, the Federal Insecticide, Fungicide, and Rodenticide Act (FIFRA) of 1947
was amended to give EPA authority to control pollution from DDT, mercury, aldrin,
toxaphene, parathion, and related chemicals. About 1 billion pounds of pesticides,
fungicides, and rodenticides are used every year in the United States. While they
contribute enormously to the success of agriculture, they can be harmful to animals,
birds and humans if not used properly.
18.14 Key Regulations 475
The Toxic Substances Control Act (TSCA) of 1976 gave EPA the authority to regulate
the development, distribution, and marketing of chemical products. Manufacturers,
importers, and processors must notify EPA within 90 days before introducing a
new chemical to the market. Certain tests (e.g., fish-kill tests) must be conducted to
determine toxicity. Approved chemicals must bear warning labels.
Many chemicals are restricted or banned under TSCA.
• The manufacture, processing and distribution of completely halogenated chloroflu-
orocarbons (CFCs) is banned, except for a small number of essential applications.
• Chromium (VI) may not be used as a corrosion inhibitor in comfort cooling towers
(CCTs) associated with air conditioning and refrigeration systems.
• Nitrosating agents may not be mixed with metalworking fluids that contain specific
substances.
• The import, manufacture, processing, or distribution of PCBs is banned unless
EPA agrees that the PCBs will be “totally enclosed.”
The Asbestos School Hazard Abatement Act (SHAA) of 1984 was passed to encour-
age the removal of asbestos from schools. In 1986, the Asbestos Hazard Emergency
Response Act was passed to correct in deficiencies in the previous Act. The final
rule, issued in 1987, required local education agencies to:
1. Inspect school buildings for asbestos-containing materials
2. Submit asbestos management plans to state governors
3. Reduce or completely eliminate all asbestos hazards.
On November 13, 1972, the Convention on the Dumping of Wastes at Sea was agreed
in London by representatives of 91 countries, including all of the world’s principle
maritime nations. The list of substances that may not be dumped includes biolog-
ical and chemical warfare agents, certain kinds of oil, certain pesticides, durable
plastics, poisonous metals and their compounds, and high-level radioactive waste.
Enforcement is left to individual countries.
476 18 Safety and Environment
In 1992, during the “Earth Summit” in Rio de Janeiro, 154 nations plus the European
community signed the United Nations Framework Convention on Climate Change
(UNFCCC). At the time, the Earth Summit was the largest-ever gathering of Heads
of State. Effective on March 21, 1994, the UNFCCC called on industrial nations to
voluntarily reduce greenhouse gas emissions to 1990 levels by the year 2000. As of
May 2015, there were 196 Parties (195 countries and 1 regional economic integration
organization) to the Framework.
In many respects, the Rio Declaration resembled the Declaration on Human Envi-
ronment issued by the Stockholm Conference in 1972. The 27 non-binding principles
of the Rio Declaration included the “polluter pays” concept and the “precautionary
principle.” The latter recommends that, before a construction project begins, an
impact study should be conducted to identify and forestall potential harm to the
environment.
The declaration asserted that present-day economic development should not
undermine the resource base of future generations. It also affirmed that industrial
nations pollute more than developing countries. (For example, on a per capita basis,
the United States emits 25 times more CO2 than India.) On the other hand, indus-
trial nations have advanced technology and greater financial resources, which enable
them to contribute more to environmental protection.
The Kyoto Protocol extended the UNFCC. Adoption committed countries to accept
that climate change was real, and that anthropogenic CO2 is the primary contributor.
The agreement was concluded in December 1997 and took force in February 2005.
As of 2015, it had been accepted by 92 countires. Significantly, Canada withdrew
its support in 2012 and the United States never adopted it. The Protocol obliges
developed countries to reduce greenhouse gas emissions while giving allowances in
transitional (developing) coiuntries. It discusses inter-government emissions trading.
The Paris Accord was signed in November 2016 by 194 members of the UNFCC.
As of December 2016, it had been ratified by 127 countries. The main difference
between the Paris Accord and previous agreements are its bottom-up approach and its
flexibility. Countries set their own targets voluntarily. Targets are not legally binding.
18.15 Climate Control: Rio, Kyoto, and Paris 477
No distinct differences are made between developed and developing economies. The
goal of the Paris Accord is to limit the increase in global average temperature to 1.5 °C
above pre-industrial levels. The NASA GISS model mentioned above predicts that
if the temperature increase reaches 2 °C, more than 150 million people could be
displaced due to famine and rising sea level. President Barrack Obama signed the
accord, but the person who followed him reneged.
Green technology does more than keep the planet clean. It also generates jobs. The
green workforce in the US rose to 8.1 million in 2015, and solar energy jobs overtook
those in the oil and gas extraction sector. Similarly, more people work in renewable
energy than in oil and gas in China [92].
18.16 Summary
Safety, reliability, and protecting the environment are inextricably linked. In modern
industry, they are viewed as prerequisites to profit. Company executives frequently
emphasize their strong commitment to worker health and safety. Their commitment
derives to a certain extent from strict legislation.
Health and safety rules address personal protection equipment, toxic substances,
equipment maintenance, worker training, and compliance monitoring. Environment
regulations fall into four main categories: air pollution (particulate matter, volatile
hydrocarbons, and harmful gases), waste water, spills and solid wastes. Harm from
chronic exposure accumulates with time.
Harm from major accidents causes tremendous short-term destruction, which
often is followed by damage that lingers for years. Many of the worst peace-time
accidents occur in coal, oil, and chemical industries. Such accidents can teach hard
lessons. But as the examples show, not everyone learns the lessons, and those who
do sometimes forget or choose to ignore them.
Things have gotten better. Before the 1970s, black lung disease, dead forests
downwind from coal-fired power plants, and pools of poison around abandoned strip
mines were just part of the price we paid for cheap power and minerals. Smoky flares
in refineries were the rule, not the exception. They “smelled like money.” More often
than not, news of a river catching fire caused laughter instead of outrage.
Today we believe that our fundamental rights include clean air, clean water, and a
safe and healthy workplace. We are products of the same social movement that created
EPA, OSHA, and similar agencies around the world. Governments are providing the
Big Stick—steep fines and possible jail time for corporate executives—but the Carrot
comes from a basic change in our fundamental values.
478 18 Safety and Environment
References
1. http://corporate.exxonmobil.com/en/company/worldwide-operations/safety-and-health/
workplace-safety. Retrieved 1 July 2015
2. http://www.shell.com/global/environment-society/safety.html. Retrieved 1 July 2015
3. http://www.kbr.com/About/Code-of-Business-Conduct/Translations/English/COBC_English.
pdf. Retrieved 1 July 2015
4. http://www.chevron.com/about/operationalexcellence/tenetsofoperation/. Retrieved 1 July
2015
5. https://en.wikipedia.org/wiki/History_of_labour_law. Retrieved 3 Feb 2015
6. https://en.wikipedia.org/wiki/Environmental_movement. Retrieved 3 Feb 2015
7. Mantoux P (2000) The industrial revolution in eighteenth century: an outline of the beginnings
of the modern factory system in England. Harper & Row. http://www.hoodfamily.info/coal/
law1775act.html. Retrieved 6 Nov 2010
8. https://en.wikipedia.org/wiki/Labor_history_of_the_United_States. Retrieved 3 Feb 2015
9. http://www.msha.gov/MSHAINFO/MSHAINF2.HTM. Retrieved 6 Mar 2015
10. https://en.wikipedia.org/wiki/1900_United_States_Census. Retrieved 3 Feb 2015
11. https://en.wikipedia.org/wiki/Child_labor_laws_in_the_United_States. Retrieved 3 Feb 2015
12. https://en.wikipedia.org/wiki/National_Child_Labor_Committee. Retrieved 3 Feb 2015
13. http://www.dol.gov/dol/aboutdol/history/flsa1938.htm. Retrieved 3 Feb 2015
14. Margulis L, Sagan D (2000) What is life? The eternal enigma. University of California Press,
Berkeley
15. Commoner B (1971) The closing circle: nature, man, and technology. Knopf, New York
16. Nobel Prize Organization. “Linus Pauling—Biographical.” www.nobelprize.org/nobel_prizes/
peace/laureates/1962/pauling-bio.html. Retrieved 26 Dec 2014
17. Carson R (1962) Silent Spring. Houghton Mifflin, New York
18. http://www.panna.org/issues/persistent-poisons/the-ddt-story. Retrieved 5 July 2015
19. http://www2.epa.gov/aboutepa/epa-history. Retrieved 3 Feb 2015
20. https://en.wikipedia.org/wiki/Environmentalism#Early_environmental_legislation. Retrieved
3 Feb 2015
21. https://en.wikipedia.org/wiki/History_of_fire_safety_legislation_in_the_United_Kingdom.
Retrieved 3 Feb 2015
22. https://en.wikipedia.org/wiki/Alkali_Act_1863. Retrieved 13 June 2015
23. https://en.wikipedia.org/wiki/Public_Health_Act_1875. Retrieved 3 Feb 2015
24. https://en.wikipedia.org/wiki/List_of_national_parks_of_the_United_States. Retrieved 3 Feb
2015
25. http://www.nps.gov/hosp/learn/historyculture/index.htm. Retrieved 3 Feb 2015
26. http://www.prod.exxonmobil.com/refiningtechnologies/
27. Couch KA, Seibert KD, Van Opdorp P (2003) Controlling FCC yields and emissions: UOP
technology for a changing environment, AM-04–45, NPRA Annual Meeting, San Antonio,
TX, 23–25 Mar 2003
28. AQMD Adopts Regulation to Reduce Particulate Emissions from Oil Refineries. AQMD Advi-
sor, 7 Nov 2003
29. Cleaner Production Initiatives—BP Kwinana Refinery (2004) Department of the Environment
and Heritage, Canberra, Australia
30. https://www.osha.gov/SLTC/hydrogensulfide/hazards.html. Retrieved 26 Jan 2016
31. U.S. Environmental Protection Agency (2004) Air trends: stratospheric ozone. National Service
Center for Environmental Publications, Cincinnati, OH
32. U.S. Environmental Protection Agency (2015) Ozone layer protection. http://www.epa.gov/
ozone/intpol. 3 Feb 2015
33. http://www.epa.gov/ozone/title6/downloads/Section_608_FactSheet2010.pdf. Retrieved 3 Feb
2015
34. http://www.epa.gov/ozone/2007stratozoneprogressreport.html. Retrieved 9 July 2015
References 479
35. Policy Implications of Greenhouse Warming (1992) Mitigation, adaptation, and the science
base. National Academy Press, Washington, DC
36. McIntyre S, McKitrick R (2003) Corrections to the Mann et. al. (1998) proxy data base and
Northern Hemispheric average temperature series. Energy Environ 14(6):751
37. Energy Information Administration, International Energy Statistics. https://www.eia.gov/beta/
international/
38. Lower SK. Carbonate equilibria in natural waters. www.chem1.com/acad/pdf/3carb.pdf
39. Andersson AJ, Yeakel KL, Bates NR, de Putron SJ (2014) Partial offsets in ocean acidification
from changing coral reef biogeochemistry. Nat Climate Change 4:56–61
40. GISTEMP Team (2015) GISS Surface Temperature Analysis (GISTEMP), NASA Goddard
Institute for Space Studies. http://data.giss.nasa.gov/gistemp/. Retrieved 26 June 2015
41. http://www.bloomberg.com/graphics/2015-whats-warming-the-world/. Retrieved 26 June
2015
42. https://en.wikipedia.org/wiki/Steam_assisted_gravity_drainage. Retrieved 6 Dec 2014
43. Gates ID, Larter SR (2014) Energy efficiency and emissions intensity of SAGD. Fuel
115:706–713
44. American Chemical Society (2016) The science and technology of hydraulic fracturing.
https://www.acs.org/content/acs/en/policy/publicpolicies/sustainability/hydraulic-fracturing-
statement.html. Accessed 3 Nov 2018
45. Galbraith K, Henry T (2013) As fracking proliferates in Texas, so do disposal wells, Texas Tri-
bune, 29 Mar 2013. http://www.texastribune.org/2013/03/29/disposal-wells-fracking-waste-
stir-water-concerns. Retrieved 26 Jan 2016
46. World Energy Interviews Barry Rosengrant (2004) Chief Executive Officer of Petromax Tech-
nologies. World Energy 7(2)
47. https://en.wikipedia.org/wiki/Davy_lamp. Retrieved 5 Feb 2015
48. https://en.wikipedia.org/wiki/Black_lung_disease. Retrieved 10 Feb 2015
49. https://en.wikipedia.org/wiki/Lakeview_Gusher. Retrieved 5 Feb 2015
50. https://en.wikipedia.org/wiki/Texas_City_Disaster. Retrieved 5 Feb 2015
51. Air Quality in the UK. Postnote, Nov 2002 (188)
52. Proceedings of the symposium: “Twenty Years after the Amoco Cadiz,” Brest, France, 15–17
Oct 1998
53. International Nuclear Safety Advisory Group (1992) The Chernobyl accident: updating of
INSAG-1, Safety Series No. 75-NSAG-7. International Atomic Energy Commission, Vienna
54. Dyatlov A (1995) Why INSAG has still got it wrong. Nuclear Engineering International, Sept
1995
55. http://www.world-nuclear.org/info/Safety-and-Security/Safety-of-Plants/Appendices/
Chernobyl-Accident—Appendix-2–Health-Impacts/. Retrieved 9 July 2015
56. Chernobyl Powers Down Permanently. CNN.com, 15 Dec 2000. http://www.cnn.com/2000/
WORLD/europe/12/15/chernobyl.shutdown/. Verified 26 Jan 2016
57. Grushevoi G (1992) World Uranium Hearing, Salzburg, pp 259–260
58. Weber U (2000) The miracle of the rhine. UNESCO Courier, June 2000
59. Tankers Have Spill-Free Year in Alaska. USA Today, 11 July 2004
60. EPA Chemical Accident Investigation Report: Tosco Avon Refinery, Martinez, California. EPA
550-R-98-009, Nov 1998
61. Robinson PR, Dolbear GE (1996) Hydrotreating and hydrocracking Fundamentals. In:
Chapter 7 in practical advances in petroleum processing. Springer, New York
62. U.S. Chemical Safety and Hazard Investigation Board (2005) Refinery explosion and fire, BP
Texas City Refinery (15 Killed, 180 Injured), Report No. 2005-04-I-TX, 23 Mar 2005
63. National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling, Report
to the President: Deepwater: The Gulf Oil Disaster the Future of Offshore Drilling, Jan 2011
64. Nandakumar A, Bousso R, Finn K (2015) BP settles 2010 Gulf oil spill claims for $18.7
billion. Reuters: DailyFinance, 2 July 2015. http://www.dailyfinance.com/2015/07/02/bp-pay-
damages-water-pollution-gulf-oil-spill/
65. https://en.wikipedia.org/wiki/Lac-Megantic_derailment. Retrieved 5 Feb 2015
480 18 Safety and Environment
Bhopal Methyl Isocyanate Leak, 424 430, 431, 433, 435, 436, 445, 446, 448,
Biomass, 182, 295–297, 345, 369, 374, 375, 450, 453, 455
377 Catalytic NOx removal, 293
Bio Oil and Petroleum Oil C0-processing, 296, Catalytic reforming
297 houdriforming, 198
Bitumen magnaforming, 198
transportation, 385–387 powerforming, 198
treating Process, 388, 389 rheniforming, 198
Black lung disease, 421, 475 ultraforming, 198
Blending, 142, 146, 155, 156, 159, 160, 172, Caustic flooding, 111, 112
190, 191, 198, 215, 217, 224, 249, 260, Cetane index, 319, 320, 467
262, 295, 297, 305, 306, 311–314, 317, Cetane number, 239, 317, 319–322, 467
323, 324, 334, 384, 386, 419, 464 C5/C6, 133, 189, 191–193
Blowout Preventer (BOP), 101, 102 C-14, 79
Boiling point, 19–23, 35, 41, 43, 57–60, 142, Chemical index, 59
144, 162, 163, 165, 169, 177, 250, 257, Chemical injection, 110, 111
266, 302, 318, 320, 406, 460 Chernobyl nuclear disaster, 425
Bonny Light, 6 Chinese, 14
BP Texas City Isomerization Unit Explosion, Chlorinated methanes, 347, 377
437 Chlorobutyl rubber, 366
Brent, 6, 7 Chloroprene, 350, 368
Bromobutyl rubber, 366 Claus process, 153, 154, 287, 290, 291
BTX Derivatives, 150, 369, 383 Cleaning Up, 317, 412, 429, 442, 457, 460
Butaldehyde, 365 Clean air acts, 422, 461, 462
Butadiene, 150, 154, 207, 214, 215, 313, 350, Clean water act, 468, 469
367, 368, 464 Climate change, 406, 407, 465, 474
Butane, 21, 41, 43, 135, 146, 150, 151, 153, Climate control, 474
155, 190, 191, 197, 199, 200, 202, 211, Cloud point, 19, 29, 239, 257, 269, 270, 321
214, 231, 234, 239, 263, 274, 299, 305, Coal, 14, 28, 40, 45, 51, 54, 70, 71, 75, 79, 94,
310, 311, 343, 352, 359, 367, 378 132, 134, 153, 231, 233, 242, 243, 294,
Butanol, 313, 365, 369 335, 345, 374, 375, 377, 396, 399, 400,
Butylenes, 150, 154, 155, 208, 215, 304, 352, 416, 461, 465, 475
359, 360 Coal mining
Butyl rubber, 366, 367, 378 explosion, 110, 397, 421
Coke formation mechanism, 225
C Coker, 135, 143, 151, 154, 173, 178, 212, 230,
Calorie value, 19, 28 233, 234, 238, 247–250, 252, 301, 320,
Canadian oil sand, 117 324, 335, 388, 419
Carbocation rearrangement, 217, 218 Cokes, 147, 247, 335, 336, 378
Carbon cycle, 67, 68, 74 Coking
Carbon isotopes, 75, 79 delayed, 143, 149, 153–155, 178, 182, 184,
Carbon monoxide, 73, 277, 345, 349, 365, 375, 212, 213, 229, 240, 242, 247–250, 252,
377, 400, 421, 424, 457, 463 294, 335, 400, 419
Casing, 98–101, 421, 440, 441 fluid, 149, 247, 249, 250
Catalysts, 26, 28, 50, 52, 53, 117, 132, 134, Cold Heavy Oil Production with Sand
142, 160–162, 174, 177–180, 183–187, (CHOPS), 119
189–191, 193–195, 197, 198, 206–209, Composition
216, 217, 219, 220, 222–231, 233, chemical, 39
235–238, 240–242, 262, 264, 274–277, elemental, 39
280, 285, 286, 288–291, 293, 294, 297, molecular, 40
302–304, 335, 343–352, 355, 357–359, Condensate, 25, 53, 63, 75, 120, 299, 300, 343,
361–363, 365, 366, 368, 370, 371, 359, 383, 384, 386, 389, 394, 409,
374–377, 387, 389, 401–404, 419, 420, 410
Index 483
Condensation, 45, 168, 172, 174–176, 180, catalytic, 136, 155, 225, 226, 257, 268,
248, 352, 359, 368, 369 269, 273, 285, 325
Conradson Carbon (CCR), 28, 136, 257, 262 Dichill, 273
Conservation, 395, 398, 399, 459, 465, 470 ketone, 270–272
Containment, 122, 234, 395, 412, 413, 429, propane, 270, 271
445 solvent, 135, 232, 257, 261, 268–270, 274,
Continuous Catalyst Regeneration (CCR), 194, 275, 369
195, 197, 225, 228, 297 Diesel
Conversion strategy, 142 additives, 322
Cracking marine, 317, 320, 322, 323
catalytic, 134, 135, 142, 146–149, 173, properties, 321
190, 194, 200, 201, 208, 211, 212, 216, Diesel engines, 98, 132, 295, 317, 319, 321,
221, 229–232, 234, 253, 254, 285, 301, 328
302, 305, 306, 310, 324, 376 Directional drilling, 102–104
fluid catalytic, 135, 151, 172, 178, 211, Dispersants, 322, 326–330, 332, 415, 423
214, 216, 220, 228, 297, 400 Dispersion agents, 415
steam, 150, 154, 211, 214, 215, 229, 253, Distillation
254, 314, 359, 367, 378 atmospheric pressure, 164
thermal, 30, 41, 55, 75, 80, 115, 134, 135, vacuum, 20–23, 135, 149, 162–164, 169,
141, 143, 149, 154, 169, 194, 200, 170, 201, 238, 247, 248, 251, 259–261,
211–213, 215, 217, 219, 229–231, 237, 400
247, 248, 251, 253, 254, 324, 337, 353, Distillation unit
358, 387 atmospheric pressure, 151, 166, 201, 315
Crack spread, 134, 136, 137 vacuum, 151, 201
Crude assay, 19, 20, 23, 31, 32, 34, 35, 39 Downstream, 123, 154, 159, 162, 177, 179,
Crude oil 180, 183, 241, 290, 294, 337, 383, 384,
extra heavy, 60, 247, 294, 385 404, 409, 424, 427, 447
heavy, 60–64 Drilling, 64, 67, 87, 90, 97–99, 102, 103, 105,
in China, 54, 55, 102, 317, 336, 349, 475 106, 117, 119, 278, 394, 417, 421, 441,
intermediate, 60, 62, 144, 159 442, 454, 455, 465
light, 60, 62, 64 Drilling Fluid (Mud), 97, 363
medium, 60, 64
sour, 27, 31, 60, 62, 401 E
sweet, 26, 27, 31, 35, 60, 81 Edwin L. Drake, 15
Cumene, 362, 369 Energy Policy Act, 409, 465
Cyclic Steam Stimulation (CSS), 113, 387, 408 Environment, 19, 41, 48, 76, 77, 87, 88, 90,
116, 131, 200, 289, 393, 395, 396, 403,
D 408, 412, 437, 439, 456, 458, 459, 462,
Deepwater Horizon Drilling Rig Blowout, 440 468, 474, 475
Dehydrogenation, 46, 172, 174, 175, 177, 180, Environmental Agencies, 457, 458
194, 195, 197, 213, 214, 217, 220, 248, Environmental laws and decrees, 397
261, 346, 350, 359, 362, 367 Environment protection, 134, 398, 457
DEMEX solvent extraction, 294, 295 Epichlorohydrin, 364
Demulsifiers, 311, 312, 322, 333 Ester base stocks
Derrick, 98, 99 aromatics, 283, 284
Desalter, 151, 161, 162, 166, 410 dibasic, 282–284, 326
Desalting, 28, 54, 142, 159–161, 163 phosphate, 286
Detergents, 155, 311, 312, 322, 326–330, 332, polyol, 278, 282–284, 326
334, 355, 357, 363, 364, 367, 368, 370, Ethylene
413, 415, 416, 419 chemicals from, 354
Dewaxed Oil (DWO), 269, 271, 276 Ethylene glycol, 346, 357
Dewaxing Ethylene oxide, 351, 357, 358
484 Index
Ethylene-Propylene-Diene (EPDM) 343, 345, 347, 352, 359, 374, 375, 379,
terpolymer, 361 383–385, 389, 391, 401, 421, 465
Ethylene-Propylene (EP) copolymer, 361 wet, 75, 249, 289, 402
Exploration, 67–69, 76, 83, 84, 88, 90, 105, Gas hydrate, 122
383, 412, 442 Gas injection, 109, 110, 112
Extreme Pressure additives, 329 Gasoline
Exxon Valdez oil spill, 428 additives, 311, 312
components, 43, 190, 200, 305, 306, 308,
F 310, 311, 313, 464
F.N. Semyenov, 14 in Europe (Petrol), 134, 137, 234, 313
Federal Water Pollution Control Act natural, 299, 352, 359
(FWPCA), 468 production, 190
Finished (Reformulated) lube oils Reformulated (RFG), 43, 155, 193, 200,
additives, 325–327 312, 313, 464
Fischer-Tropsch process, 277, 346, 374 Gasoline engines, 132, 189, 200, 306, 328
Flash point, 19, 29, 239, 257, 266, 280, 311, Gas recovery, 97, 304
314–316, 318, 321, 323, 460, 467 Gas-to-Liquid (GTL) process, 277, 375
Flexicoking, 240, 247, 250, 251 Geological time scale
Fluid Catalytic Cracker (FCC) age, 72, 76, 79, 94
processes, 173, 216, 221, 225, 234, 335 eon, 72, 73
Fluorecence, 30 era, 73, 74
Formaldehyde, 313, 328, 346, 347, 350, 351, Geophysical methods for exploration, 83, 87
377, 464 Glycerol, 363, 364
Formation, 29, 52, 54, 67, 69, 70, 76, 77, 79, Grease, 138, 154, 256, 265, 278, 281, 302, 315,
84, 87–89, 93, 97, 103, 104, 110, 111, 325, 326, 330, 332, 334, 336, 338, 460
113, 117, 121, 122, 132, 174, 180, 190, Greenhouse gases, 118, 122, 295, 296, 388,
197, 203, 214, 216, 225, 232, 233, 242, 406, 408, 409, 474
247, 249, 252, 254, 257, 261, 265, 267, Gulf war intentional oil spill and oil well fires,
269, 276, 283, 294, 299–301, 312, 322, 429, 454
327, 330–332, 336, 349, 356, 357, 367,
372, 375, 383, 387, 391, 423, 440 H
Fossil fuels, 71, 294, 296, 406, 407, 465 Heating oils, 20, 21, 143, 145, 146, 154, 165,
Fossil hydrocarbons, 70, 80, 395 173, 211, 302, 320, 323
Freeze point, 19, 29, 239, 316 Heavy oil
Friction modifier, 326, 327, 330 transportation, 385, 386
Fuel oils, 21, 134, 135, 137, 143, 145, 146, Heteroatom-containing hydrocarbons, 39, 41,
153, 154, 156, 172, 173, 178, 211, 215, 47, 55, 255
220, 240, 251, 253, 288, 301–304, Horizontal drilling, 101–103
322–324, 343, 348 Huff and Puff, 113, 116
Furfural extraction, 267 Hydraulic fracturing (Fracking), 120, 121, 299,
383, 408
G Hydride transfer, 198, 219, 225
Gas Hydrocracker
acid, 177, 289, 292, 293, 300, 384, 389, slurry-phase, 143, 154, 173, 178, 233
391 Hydrocracking
dry, 75 ebulalted-bed catalytic, 143, 153, 154, 173,
natural, 27, 41, 48, 53, 54, 69, 75, 79–81, 178, 228
87, 94, 103, 106, 109, 110, 115, 120, fixed-bed catalytic, 172, 196, 198, 233,
121, 123, 129, 141, 155, 184, 215, 251, 235, 430
277, 285, 287, 289, 292, 294, 299–305, slurry-phase, 154, 231, 233, 242–243, 254
Index 485
Hydrodemetallation (HDM), 142, 177, 183, 238, 239, 247, 301–304, 314–316,
184, 223, 285 323–325, 343, 386, 389, 416
Hydrodenitrogenation (HDN), 142, 177, Kyoto Protocol, 458, 461, 474
182–184, 285, 431, 434
Hydrodeoxygenation (HDO), 177, 178, 182, L
285, 296, 297 Labor Legislation and Decrees, 396
Hydrodesulfurization (HDS), 142, 177, 180, Lac-Megantic Oil Train Derailment, 443
181, 184, 285 Lakeview blowout, 454
Hydroformylation (Oxo Process), 368 Liquefied Petroleum Gas (LPG), 110, 143, 146,
Hydrogen 154, 155, 164, 170, 173, 178, 186, 211,
from methane, 343 214, 215, 217, 223, 230, 275, 302, 303,
production, 142, 183, 285, 287, 304 305, 346, 359
purification, 286, 288 London killer fog, 454
Hydrogenation, 45, 46, 135, 172, 175, 194, Lube assay, 260
198, 208, 211, 224, 232–234, 261, 275, Lube Plant Feedstocks, 258, 259
295, 296, 346, 347, 368, 370, 376, 377 Lube processing, 259, 264
Hydrogen cyanide, 351, 365, 444 Lubricant
Hydrogen sulfide, 27, 48, 53, 72, 123, 185, base Oil Properties, 254
240, 249, 287, 289, 290, 299, 347, 363, base Oils, 253
401, 432 basestock API Categories, 256
Hydrotreating
catalysts, 142, 174, 178–180, 184, 185, M
228, 237, 297, 431, 433 Management of Change (MOC), 453
chemistry, 180 Marine Protection, research and Sanctuaries
process flow, 178 Act, 469
Material Safety Data Sheet (MSDS), 459, 460
I Mercaptan, 27, 48, 49, 55, 135, 142, 186, 208,
Ignacy Lukasiewicz, 14 291, 299, 363
In-situ combustion, 114 Mercaptan Oxidation (Merox), 27, 48, 135,
Integrated Gasification Combined Cycle 151, 185, 186
(IGCC), 377 Metal-containing hydrocarbons, 52, 55
Isodewaxing, 232, 261, 264, 275, 285, 325 Metal content, 19, 28, 35, 40, 56, 240, 335
Isomerization Metal deactivator, 311, 312, 316, 322, 327, 332
butane, 135, 153, 189, 190 Methanation, 286, 345
C5/C6, 189, 191–193 Methane
Isopropyl Alcohol (IPA), 314, 362, 466 chemicals from, 343, 344
Isotope ratio mass spectrometry, 79, 80 Methane to methanol, 345
Methanol to Gasoline (MTG), 345, 346
J Methanol to Olefins (MTO), 345, 346
James Oakes, 11, 14 Methyl Ethyl Ketone (MEK), 269, 325, 369
James Young, 14 Michael Dietz, 14
Jet fuel, 20, 21, 29, 40, 134, 143, 144, 146, Microbial stimulation, 112
147, 153, 155, 163, 165, 167, 170, 172, Micro-Carbon Residue (MCR), 28, 257
173, 178, 211, 302–304, 314–316, 320, Midstream, 123, 383, 384, 408
323–325, 343, 346, 430 Migration
primary, 91, 92
K secondary, 91, 92
Kerogen, 19, 29, 69, 74–79, 91, 116, 117, 211, tertiary, 91
294, 389, 419
Kerogen type, 77 N
Kerosene, 14, 20, 21, 29, 40, 129–131, 134, Naphtha, 20, 21, 35, 40, 45, 129, 134–136,
135, 137, 143, 146, 151, 153, 155, 138, 143, 146, 150, 151, 153, 163–168,
163–167, 170–173, 178, 180, 211, 230, 170, 172, 173, 177–180, 184, 185, 189,
486 Index
190, 194, 195, 198, 213–217, 220, 224, Oxidative resistance, 257, 281
227, 230, 234, 238, 239, 249, 251, 274, Oxygen-containing hydrocarbons, 51
285, 294, 301, 302, 310, 314, 315, 320, Ozone
324, 352, 359, 374, 378, 383, 384, 386, stratospheric, 405
388, 389
Naphthenes (Cycloparaffins), 44, 274 P
Natural forces, 106, 109, 412 Paleotransformation
Natural gas catagenesis, 75
liquefaction, 301, 384, 389, 390 diagenesis, 74, 75
Natural Gas Liquid (NGL), 299, 302, 303, 352, metagenesis, 75
359 Paris accord, 474, 475
Near-misses, 420, 445, 451 Partial oxidation, 234, 285, 286, 345, 352, 374,
Nitric acid, 327, 348, 370, 372 375
Nitrogen, 6 Particulate Matter, 27, 399, 400, 402, 475
Nitrogen-containing hydrocarbons, 49 PeeDee Belemnite (PDB) standard, 79
Nitrogen content, 19, 28, 31, 35, 39, 49, 55, Petrochemical plants, 132, 138, 148, 150, 151,
240, 297, 349, 404 206, 289, 299, 378, 404
Nitrogen Oxides (NOx), 27, 293, 317, 400, Petrochemicals, 21, 132, 135, 145–148, 193,
404, 464 207, 255, 279, 280, 302, 305, 337, 343,
NMP Extraction, 265, 266 344, 352, 359, 369, 378–380, 383, 399,
Non-dispersion methods, 415 400, 419
Nonylphenols, 363, 370 Petroleum
Nylons, 348, 357, 362, 367, 370, 371 accumulation, 84, 91, 93
alternatives, 294, 295, 318
O bulk properties, 19, 31, 35
Occupational Safety and Health Administration characterization, 58, 60
(OSHA), 456, 457, 459, 460, 475 classification, 57, 60, 61
Octane Number (Rating) composition, 39, 63
Motor (MON), 198, 309 gas, 21, 151, 154, 197, 297, 304
Research (RON), 194, 198, 309 geochemistry, 90
Oil Creek, 15 historical events, 129
Oil sand molecular types, 257, 320
cleaning up, 317, 429, 442, 457, 460 origin, 71, 75
transportation, 387 processing, 129, 394, 417
upgrading, 387 products, 123, 134, 137, 144, 172, 302,
Olefins, 40, 41, 53, 55, 135, 138, 143, 146, 303, 323, 338
148, 150, 153–155, 173, 174, 178, 180, system, 67, 69, 70, 84
190, 191, 193, 194, 198, 200, 202, 203, Petrolia, 15
205, 206, 208, 211–217, 219, 220, 224, Phenol extraction, 267, 268
225, 230, 232, 234–238, 249, 254, 264, Physical removal, 413
275, 276, 281, 285, 301, 305, 310, 314, Phytane, 41
316, 324, 326, 331, 337, 343, 346, 352, Play, 67, 75, 84, 102, 180, 379, 396, 406, 412
355, 359, 363, 365, 367, 368, 374, Pollution control and abatement technology,
376–378, 383, 466 399
Open pit mining, 117 Polyalphaolefins (PAO), 279, 281, 283, 284,
Optical activities, 19, 30, 71 326
Organization of Petroleum Exporting Countries Polybutylenes, 365, 366
(OPEC), 129 Polyalkylene Glycol (PAG), 283
Oswald process, 348 Polyethylenes
Otto cycle, 307, 308 High Density (HDPE), 354
Outcrop, 67, 84, 86 Low Density (LDPE), 354
Overburden, 69, 75 mechanism, 355, 356
Index 487
Polyethylenes Glycol (PEG), 332, 357, 358 integrated, 138, 145, 148, 150, 231, 296
Polyisobutylene (PIB), 286, 329, 366 processes, 57, 152, 160, 305, 306, 320, 338,
Polymer flooding, 111, 112 373
Polymerization, 53, 134, 135, 142, 149, 152, products, 338
153, 180, 189, 198–200, 207–209, 216, topping (Simple), 144–146
220, 248, 302, 306, 310, 324, 327, 346, Refractive index, 19, 30, 260
353, 355, 358, 361, 366, 370, 460 Refuse act, 467, 468
Polypropylenes, 155, 208, 214, 305, 332, 349, Regulations, 122, 123, 159, 216, 291, 305, 310,
357–359, 361, 365, 378, 415 313, 317, 322, 338, 395–397, 399, 425,
Polyvinyl Acetate (PVA), 358 429, 442, 444, 456, 457, 459, 461, 464,
Polyvinyl Chloride (PVC), 349, 357, 358, 413 469, 471, 475
Pour point, 19, 28, 29, 35, 36, 135, 136, 251, Reid Vapor Pressure (RVP), 30, 31, 146, 200,
257, 261, 269, 271, 272, 274, 275, 296, 310, 311, 313, 464
279–281, 283, 284, 286, 321, 323, 325, Renewable fuel mandates, 465
327, 333, 386 Reservoir, 19, 25, 57, 63, 64, 67–69, 75, 76,
Pour point depressants, 333 79, 83, 84, 86, 88, 90, 92, 97, 98, 103,
Pressure Swing Adsorption (PSA), 237, 238, 106, 107, 109–111, 113, 115–117,
286–288, 345 119–123, 184, 294, 393, 440, 442
Pristane, 41 Reservoir rock, 69, 84, 86, 91, 99, 101, 110,
Production, 25, 26, 29, 48, 51, 53, 57, 63, 64, 121, 408
68, 69, 76, 83, 93, 94, 97, 101, 103–107, Resid Fluid Catalytic Cracker (RFCC), 143,
109–111, 113–115, 119, 123, 129, 131, 146, 153, 154, 173, 178, 223
132, 138, 145, 153, 160, 172, 174, 182, Resource Conservation and Recovery Act
185, 198, 199, 211, 215, 216, 222, 226, (RCRA), 419, 459, 470, 471
228, 230–232, 234, 236, 238, 247, 250, Rio earth summit, 474
253–255, 261, 263, 268, 275, 294, 296, River and Harbor Act, 468
299, 337, 343, 346, 348–352, 356, 357, Rockefeller, John D., 131
363, 365, 367, 374, 375, 378, 379, 383,
384, 388, 389, 393–395, 400, 401, 405, S
408, 412, 437, 442, 443, 462, 465, 466 Safe Drinking Water Act, 122, 409, 465, 469
Propane, 41, 43, 144, 151, 155, 173, 197, 200, Safety, 31, 48, 83, 101, 123, 206, 257, 267,
202, 203, 208, 214, 231, 261–263, 394–397, 399, 421, 424, 425, 439, 442,
269–272, 274, 294, 299, 302–305, 325, 444, 445, 447, 452, 454–457, 459, 470,
343, 352, 359, 391, 417, 432, 465 472, 475
Propylenes, 146, 150, 154, 155, 208, 214, 215, Sandoz chemical spill and fire, 427
222, 227, 231, 253, 279, 281, 302–305, SARA analysis, 62, 63
331, 346, 352, 359–365, 369, 378 Saturation
Prospect, 67, 68, 84, 87 aromatics, 45, 172, 174, 177, 180, 182,
234, 236, 288
R olefins, 177, 180
Ratawi, 6, 7 Seismic exploration, 84, 87
Recovery Significant accidents, 420
heavy oil, 116 Simulated Distillation (SimDist), 23, 24, 257
primary, 106, 107, 109 Sludge, 198, 199, 257, 261, 265, 283, 317,
secondary, 109, 110 328, 331, 394, 410, 411, 417–419
tertiary, 109, 110 Smoke point, 19, 29, 239, 314–316
Refineries Solid waste
capacities, 132, 138–140 disposal, 417, 470
categories, 144 recovery, 417
configuratiion, 150, 159 Solid waste handling, 394
conversion, 145–148, 150, 152, 154, 172, Solvent deasphalting, 243, 261–263, 336
417 Solvent extraction, 135, 142, 146, 240, 261,
hydroskimming, 144–147 265, 268, 325, 394, 411
488 Index
Solving refining, 142, 173, 255, 258, 261 Tianjin storage station explosion, 444
Specialty oils/products, 256, 269, 278, 325, Titusville, Pennsylvania, 15
334 Toe-to-Heel Air Injection (THAI), 115, 387
Specific gravity, 20, 25, 31, 35, 49, 58–60, 116, Toluene
220, 239, 266, 334 derivative, 372
Spent catalysts, 185, 222, 241, 242, 394, 404, disproportionation, 373
417, 418, 420 Toluene Diisocyanates (TDI), 372
Standard oil company, 131 Tosco avon hydrocracker deadly pipe rupture,
Steam Assisted Gravity Drainage (SAGD), 430
119, 120, 123, 383, 387, 408 Total Acid Number (TAN), 20, 28, 160
Steam flooding, 88, 113, 114, 119 Toxic Substances Control Act (TSCA), 473
Steam Methane Reforming (SMR), 234, 236, Transportation, 27, 29, 48, 53, 117, 123, 124,
237, 285, 286, 304, 305, 343, 347, 377 132, 138, 141, 145, 166, 170, 182, 183,
Storage, 39, 40, 52, 86, 106, 119, 123, 132, 189, 295, 300, 306, 313, 317, 372,
142, 150, 159, 182, 183, 186, 257, 312, 383–385, 389, 417, 443, 455, 466, 470,
313, 322, 323, 332, 383–385, 391, 471
399–401, 403, 404, 409, 412, 413, 417, Trap
419, 422, 424, 429, 438, 460, 464, 470, anticline, 84, 85, 93
471 fault, 84
Stratum, 67 salt-dome, 84, 85
Sulfide oxides, 80, 216, 293, 399, 401, 402 stratigraphic, 86, 90
Sulfur-containing hydrocarbons, 27, 47, 49 Treating, 120, 142, 146, 161, 177, 199, 289,
Sulfur content, 19, 20, 26, 31, 34, 35, 49, 60, 291, 410, 434
81, 160, 170, 181, 194, 240, 310, 311, Treatment
313, 314, 317, 321, 323, 465 primary, 394, 410, 411
Sulfur incorporation, 80 secondary, 411, 469
Sulfur recovery, 142, 224, 290–293, 300, 336, tertiary, 394, 412
393, 400 Trinitrotoluene (TNT), 372, 422, 444
Sulfur removal, 170, 180, 182, 287, 293, 297 True Boiling Point (TBP), 20
Supercritical fluid extraction, 417 2-Ethylhexyl Alcohol, 365
Superfund, 457, 472
Surfactant (Micellar) flooding, 111 U
Sweetening, 135, 142, 146, 177, 288, 289, 300 Ultra-Low Sulfur Diesel (ULSD), 181, 182,
Syngas to Gasoline Plus (STG+), 376 317, 321
Syngas to methanol, 377 Upstream, 76, 123, 223, 383, 384, 387, 403,
Synthesis gas (Syngas), 247, 250, 277, 285, 408
286, 335, 343, 345, 374, 378, 383 Urea, 63, 293, 346–351
Synthetic crude oil (Syn crude), 116, 117, 247, US Environmental Protection Agency (EPA),
384, 388, 389 45, 312, 317, 429
Synthetic lubricant oils, 278
V
T Vacuum gas oil, 21, 22, 52, 143, 151, 164, 169,
Tackiness agents, 327, 332 172, 173, 178, 182, 222, 227, 233, 234,
Tank farm, 142, 150, 159, 384, 385 239, 259, 262, 294, 430
Tapis, 6 Vacuum residue, 6, 22, 23, 35, 143, 151, 153,
Tar, 26, 40, 116, 135, 153, 194, 203, 242, 336, 164, 170, 177, 178, 223, 233, 240, 242,
387, 419, 430 243, 249, 261, 262, 294, 335, 336, 345
Texas city ammonium nitrate explosion, 454 Vinyl acetate, 350, 351, 356, 358
Thermal cracking, 30, 41, 55, 75, 80, 115, 134, Vinyl chloride, 349–351, 358, 464
135, 141, 143, 149, 154, 169, 194, 200, Visbreaking
211, 213, 215, 217, 219, 229–231, 237, coiled, 252
247, 248, 251, 253, 254, 324, 337, 353, soaker, 253
358, 387 Viscosity
Thermal recovery, 110, 113, 387 dynamic, 256
Index 489
kinematic, 25, 36, 256, 260, 261, 280, 281, Wax, 20, 29, 43, 54, 57, 135, 138, 142, 154,
321 213, 255–258, 260, 261, 264, 268–271,
Viscosity Index (VI), 26, 45, 258, 279, 325 273–278, 302, 303, 325, 333, 335, 338,
Viscosity index improvement, 260, 261 343, 375
Viscosity index improvers, 327, 331 Wax hydrocracking, 274, 276
Volatile Organic Carbon (VOC), 313, 404, Wax Hydroisomerization, 255, 258, 261, 269,
457, 464 274, 275, 325
Volatility, 20, 30, 184, 206, 255, 257, 267, Well completion, 100, 101, 441, 455
278, 279, 283, 310, 314, 315, 377, Well logging, 87, 88
466 Wellsite pretreatment, 123
Western Canadian Sedimentary Basin, 94, 95,
W 117
Waste water from petroleum processing, 408 Whale oil, 131
Waste water from petroleum production, 408
Waste water treatment, 394, 400, 410 X
Water gas shift, 285, 375 Xylene derivatives, 373
Water Quality Act, 468, 469 Xylenes, 146, 148, 150, 215, 279, 337, 369,
Watson Characterization Factors (Kw), 58, 60 373, 374, 378, 383