You are on page 1of 40

16-1

Gas processing covers a broad range of operations to pre-


pare natural gas for market. Processes for removal of contami-
nants such as H
2
S, CO
2
and water are covered extensively in
other sections of the Data Book. This chapter will cover the pro-
cesses involved in recovering light hydrocarbon liquids either
for sale when their value as liquids is higher than their value as
gas components or they must be removed to avoid condensation.
The equipment components included in the processes described
are covered in other sections of the Data Book. This section will
bring those components together in process configurations used
for liquid production.
INTRODUCTION
The recovery of light hydrocarbon liquids from natural gas
streams can range from simple dew point control to avoid liquid
formation to deep ethane extraction. The desired degree of liq-
uid recovery has a profound effect on process selection, com-
plexity, and cost of the processing facility.
The term NGL (natural gas liquids ) is a general term which
applies to liquids recovered from natural gas and as such refers
to ethane and heavier products. The term LPG (liquefied petro-
leum gas) describes hydrocarbon mixtures in which the main
components are propane, iso and normal butane, propene and
butenes. Typically in natural gas production olefins are not
present in LPG.
Typically, modern gas processing facilities produce a single
ethane plus product (normally called Y-grade) which is often
sent offsite for fractionation and processing. Whether accom-
plished on-site or at another facility, the mixed product will
typically be fractionated to make products such as purity eth-
ane, ethane-propane (EP), commercial propane, isobutane, nor-
mal butane, mixed butanes, butane-gasoline (BG), and gasoline
(or stabilized condensate). The degree of fractionation and the
liquid products is market and geographically dependent.
Early efforts in the 20th century for liquid recovery involved
compression and cooling of the gas stream and stabilization of a
gasoline product. The lean oil absorption process was developed
in the 1920s to increase recovery of gasoline and produce products
with increasing quantities of butane. These gasoline products
were, and still are, sold on a Reid vapor pressure (RVP) specifica-
tion. Vapor pressures such as 69 or 83 kPa (abs) are common
specifications for gasoline products. To further increase produc-
tion of liquids, refrigerated lean oil absorption was developed in
the 1950s. By cooling the oil and the gas with refrigeration, the
absorber vapor outlet is leaner and propane product can be recov-
ered. With the production of propane from lean oil plants, a mar-
ket developed for LPG as a portable liquid fuel.
In lieu of using lean oil, refrigeration of the gas can be used
for propane and heavier component recovery. The use of straight
refrigeration typically results in a much more economical pro-
cessing facility than using lean oil. The chilling of the gas can be
accomplished with mechanical refrigeration, absorption refrig-
eration, expansion through a J-T valve, or a combination. In
order to achieve still lower processing temperatures, cascade
refrigeration, mixed refrigerants, and most significantly turbo-
expander technologies have been developed and applied. With
these technologies, recoveries of liquids can be significantly in-
creased to achieve deep ethane recoveries. Early ethane recov-
ery facilities targeted about 50% ethane recovery. As processes
developed, ethane recovery efficiencies have increased to well
over 90% in well integrated facilities.
In some instances heavy hydrocarbons are removed to con-
trol the hydrocarbon dew point of the gas and prevent liquid
from condensing in pipeline transmission and fuel systems. In
this case the liquids are a byproduct of the processing and if no
market exists for the liquids, they may be used as fuel. Alterna-
tively, the liquids may be stabilized and marketed as conden-
sate.
GAS COMPOSITION
The gas composition has a major impact on the economics of
NGL recovery and the process selection. In general, gas with a
greater quantity of liquefiable hydrocarbons produces a greater
quantity of products and hence greater revenues for the gas
processing facility. Richer gas also entails larger refrigeration
duties, larger heat exchange surfaces and higher capital cost for
a given recovery efficiency. Leaner gases generally require more
severe processing conditions (lower temperatures) to achieve
high recovery efficiencies and incur a higher cost per unit of
liquid product.
Gases are typically characterized by the cubic meters of re-
coverable hydrocarbons per thousand cubic meters of gas. This
is commonly expressed as liquid content. Liquid content was
traditionally meant to apply to propane and heavier compo-
nents but is often used to include ethane. The liquid content of
a gas can be calculated as shown in Example 16-1.
The other major consideration in the evaluation of NGL re-
covery options is the specification of the residue sales gas. Sales
specifications are usually concerned with a minimum Higher
Heating Value (HHV) of the gas, but in some instances the
maximum HHV can also be a consideration. The calculation of
HHV is covered in Section 23 and in more detail in GPA Stan-
dard 2172, Calculation of Gross Heating Value, Relative Den-
sity, and Compressibility Factor for Natural Gas Mixtures from
Compositional Analysis. In addition, for some gas sales the
maximum and mininum Wobbe Number of the gas may be spec-
ified. For more information on the calculation of Wobbe Num-
ber, See Section 1 definitions.
Removal of liquids results in gas shrinkage and reduction
of the HHV. This shrinkage represents a loss of revenue for the
gas sales which must be considered in the economics of an NGL
recovery plant. In general, sales gas specifications set the mini-
mum HHV at 35.4-37.3 MJ/m
3
. Thus, if any components such as
nitrogen or CO
2
are present in the gas, sufficient ethane and
heavier components must remain in the gas to meet the heating
value specification. If little nitrogen or CO
2
is present in the
gas, the recovery level of the ethane and heavier components is
then limited by markets, cost of recovery, and gas value. The
SECTION 16
Hydrocarbon Recovery
16-2
calculation of HHV and shrinkage cost is illustrated in Example
16-1.
Example 16-1 Find the liquid content of the gas mixture in
Fig. 16-1. Find the HHV of the feed gas and the HHV of the
residue gas with the following NGL recovery efficiencies: C
2

90%, C
3
98%, iC
4
/nC
4
99%, C
5
+ 100%. What is the shrink-
age cost at $3.80/MkJ?
Solution Steps:
Solution is shown in Fig. 16-1.
From Fig. 23-2 obtain the m
3
/kmol for each of the compo-
nents. Multiply the mole fraction of each component (mole
%/100) by the total moles of inlet gas and divide by 0.02369
km
3
/kmol to get the m
3
liq/1000 m
3
gas of each component. The
total m
3
liq/1000 m
3
gas from Fig. 16-1 is 0.4169.
For the recoveries specified the net m
3
/day and residue com-
position can be found as shown in Fig. 16-1. In order to compute
the HHV of the two streams, the HHVs of each component are
found in Fig. 23-2. Multiplying the individual HHVs by the
mole % gives a total HHV of 41.628 for the feed gas and 36.260
for the residue gas.
The shrinkage volume can be found by the difference of the
volume of the feed gas times the HHV and the volume of residue
gas times its HHV. This volume is then multiplied by $3.80/
MkJ to get the shrinkage value of the NGLs.
Shrinkage Value = [(9.34 41.628) (8.38 36.260)]
$3.80/MkJ = $322 800/day
The value of the NGLs in $/m
3
versus the value of the com-
ponents in the residue gas in $/m
3
or the spread between these
values is the primary economic criteria for NGL recovery proj-
ect evaluations.
LIQUID CONTENT CALCULATION
Component
Feed Gas
Mole %
m
3
/kmol
Available
Estimated
Recovery %
Net m
3
/day
Residue Gas
Mole %
m
3
/1000 m
3
m
3
/day
N
2
1.000 1.115
CO
2
3.000 3.346
C
1
85.000 94.808
C
2
5.800 0.08445 0.2077 1936 90 1742 0.647
C
3
3.000 0.08686 0.1102 1030 98 1009 0.067
IC
4
0.700 0.10322 0.0306 285 99 283 0.008
NC
4
0.800 0.09950 0.0337 315 99 311 0.009
IC
5
0.300 0.11552 0.0147 137 100 137 0.000
NC
5
0.200 0.11433 0.0097 90 100 90 0.000
C
6
+ (Set as nC
6

for example)
0.200 0.12980 0.0110 103 100 103 0.000
Total 100.000 0.4169 3896 3675 100.000
MSm
3
/day 9.34 8.38
SHRINKAGE CALCULATION
Component
Feed Gas
Mole%
Residue
Gas Mole %
HHV MJ/m
3
Feed Gas
MJ/m
3
Residue Gas
MJ/m
3
N
2
1.000 1.115 0.0 0.00 0.00
CO
2
3.000 3.346 0.0 0.00 0.00
C
1
85.000 94.808 37.707 32.051 35.749
C
2
5.800 0.647 66.067 3.832 0.427
C
3
3.000 0.067 93.936 2.818 0.063
IC
4
0.700 0.008 121.404 0.850 0.010
NC
4
0.800 0.009 121.792 0.974 0.011
IC
5
0.300 0.000 149.363 0.448 0.000
NC
5
0.200 0.000 149.656 0.299 0.000
C
6
+ 0.200 0.000 177.554 0.355 0.000
Total 100.000 100.000 41.628 36.260
MSm
3
/day 9.34 8.38
FIG. 16-1
Solution to Example 16-1
16-3
DEW POINT CONTROL
Retrograde condensation has long been known to occur at
reservoir conditions. Recognition that it also occurs in typical
processing conditions was an early result of computer calcula-
tions using equations of state to predict vapor-liquid behavior.
The phenomenon is illustrated in Fig. 16-2 showing dew point
calculations for a gas stream leaving a separator at 38C and
7000 kPa (abs). These dew point curves show that as the pres-
sure is reduced, liquid is formed. The heavier the hydrocarbon,
the more the dew point temperature increases as the pressure is
lowered. The cricondentherm of the dew point curve is primarily
determined by the presence of the heaviest component in the
gas rather than just the total quantity of the heavy component
in the feed gas so accurate determination of the heaviest com-
ponents is important in establishing the phase envelope.
When gas is transported in pipelines, consideration must be
given to the control of the formation of hydrocarbon liquids in
the pipeline system. Condensation of liquid is a problem in me-
tering, pressure drop and safe operation. Condensation of liquid
can also be a major problem due to two-phase flow and liquid
slugging.
To prevent the formation of liquids in the system, it is neces-
sary to control the hydrocarbon dew point below the pipeline
operating conditions. Since the pipeline operating conditions
are usually fixed by design and environmental considerations,
single-phase flow can only be assured by removal of the heavier
hydrocarbons from the gas.
Low Temperature Separation
Several methods can be used to reduce the hydrocarbon dew
point. If sufficient pressure is available, the removal can be ac-
complished by expansion refrigeration in an LTS unit. The ex-
pansion refrigeration system uses the Joule-Thomson effect to
reduce the gas temperature upon adiabatic expansion. This
temperature reduction results in not only hydrocarbon liquid
condensation but also water condensation. The water is gener-
ally removed as hydrates in this process, melted and removed.
Thus, the process can actually accomplish dew point control of
both water and hydrocarbon in a single unit.
Fig. 16-3 shows one example of an LTS system. The high
pressure gas may first go through a heater but this is often not
needed, depending on the gas conditions. The gas then enters
the heat exchanger coil in the bottom of the separator where the
gas is cooled by exchange with the condensed liquid and by
melting the hydrates formed. Any water or condensate pro-
duced at this point is removed in the high pressure separator.
The gas from the separator is then heat exchanged with the
FIG. 16-2
Typical Low Pressure Retrograde Condensation Dewpoint Curves
3
16-4
outlet product gas for further cooling. The temperature must be
controlled at this point to prevent hydrate formation in the ex-
changer. The gas from this point passes through the pressure
reducing valve where the Joule-Thomson expansion occurs. The
hydrocarbon liquid and hydrates produced from this expansion
fall to the bottom of the low temperature separator. The hy-
drates are melted and both the water and condensate are re-
moved by level control. The gas leaving the separator has a hy-
drocarbon dew point equal to the temperature and pressure of
the separator.
The hydrocarbon and water dew points achievable with this
process are limited by the pressure differential available as well
as the composition of the feed gas. The LTS system can only be
used where sufficient pressure is available to perform the de-
sired processing and separation. It is an attractive process step
if sufficient liquid removal can be achieved at the available op-
erating conditions. A further modification to this process is to
add glycol injection to the high pressure gas to allow the achieve-
ment of lower water dew points when available pressure is lim-
ited. Fig. 16-4 shows an LTS system with glycol injection. The
use of the glycol eliminates the need to heat the LTS liquid
phase and helps to ensure that no hydrate formation will block
the process equipment upstream of the LTS.
The LTS may not work effectively if the removal of heavy
hydrocarbon then changes the phase envelope so that partial
condensation becomes infeasible. This is typical of lean gas
containing small amounts of heavy hydrocarbons.
Refrigeration
Often excess pressure is not available to operate an LTS sys-
tem. An alternative to the expansion refrigeration system is to
utilize a mechanical refrigeration system to remove heavy hy-
drocarbon components and reduce the gas dew point. The sche-
matic for a refrigeration dew point control unit is shown in Fig.
16-5. This process flowsheet is essentially the same as that used
for straight refrigeration NGL recovery. The gas pressure is
generally maintained through the process allowing for equip-
ment pressure drops. The gas is heat exchanged and then cooled
by the refrigeration chiller to a specified temperature. Liquid is
separated in the cold separator. The temperature of the separa-
tor is set to provide the desired dew point margin for sales gas
operations. This temperature specification must take into ac-
count the gas which is recombined from the liquid stabilization
step as well as potential variations in the feed gas pressure.
Provision must be made in this process for hydrate preven-
tion. This can be accomplished by either dehydration upstream
of the unit or by integrating the dehydration with the refrigera-
tion unit. Use of glycol injection is usually the most cost effec-
tive means of controlling water dew points. The only drawback
is that the refrigeration must be in operation to accomplish the
dehydration. If it is desired to operate the dehydration at times
independent of the refrigeration, then separate units are used.
FIG. 16-3
Low-Temperature Separation Unit
4
16-5
FIG. 16-4
Low-Temperature Separation System with Glycol Injection and Condensate Stabilization
4
FIG. 16-5
Straight Refrigeration Process
16-6
Stabilization
One of the problems in using dew point control units of both
expansion LTS and mechanical refrigeration systems is the dis-
position of the liquids removed. The liquids must be stabilized
by flashing to lower pressure at high temperature or by using a
stabilization column. When the condensate is flashed to a lower
pressure, light hydrocarbons are liberated which can often be
used in a fuel gas system.
The stabilization column can produce a higher quality and
better controlled product. The condensate stabilizer is usually
a top feed column which runs at a reduced pressure from the
cold separator and has a reboiler to produce a specified vapor
pressure product. The overhead vapor is either sent to fuel as
shown in Fig. 16-4 or recompressed and combined with the
sales gas as shown in Fig. 16-5. The column contains either
trays or packing to provide necessary mass transfer for stabili-
zation of the liquid feed. After stabilization, the product is
cooled and sent to storage.
Compact Separation Designs
Compact process configurations have been commercialized
to take advantage of gas expansion for liquid separation. Each
of these processes use static equipment to achieve the desired
separation and are focused on replacing Joule-Thomson expan-
sion valves and/or turbo-expanders.
One of these processes which is currently used in several
operations is the Twister

technology. This process (Fig. 16-6)


uses a supersonic nozzle in which the pressure is reduced via
isentropic expansion such that liquid is formed. Saturated feed
gas is passed across a vane ring which induces a swirling mo-
tion where after the swirling gas is expanded in a Laval-type
nozzle causing pressure and temperature to drop while the flow
becomes supersonic. The centrifugal motion forces the liquid
droplets to agglomerate at the wall from where it is drained
from the apparatus. The dry vapor stream is then decelerated
in a diffuser duct to pipeline velocity, causing the pressure to
increase to 7080% of the feed pressure. Ample testing has
shown that the Twister process has a high isentropic efficiency
while separating almost all of the condensed liquid.
This technology is used for water and hydrocarbon dewpoint
control in both onshore and offshore locations, especially where
space and mass are at a premium. Since the expansion cooling
in Twister is close to isentropic, more valuable NGL liquids are
produced compared to a JT LTS process. A Twister based dew
pointing process can often operate without glycol or methanol
systems for hydrate suppression.
Another process uses a vortex tube device to affect the sepa-
ration. The vortex tube is based on the Ranque-Hilsch tubes de-
veloped in the 1940s. These tubes have been used as laboratory
devices and small scale coolers. The working principle of these
devices is the same. (Fig. 16-7) A gas is injected tangentially
through a nozzle into the center of the tube where it expands to
a low pressure. The gas flows cyclonic to the far end of the tube.
During this flow, two temperature zones are formed, a warm
zone near the wall and a cooler zone near the center. At the end
of the tube the center gas is deflected and returns along the tube
through an orifice near the inlet nozzle. The tube is therefore
capable of producing two gas streams at different temperatures.
The cold gas is at a temperature below that achievable with an
isenthalpic expansion. If the two outlet streams were to be mixed,
the combined temperature would be equal to the temperature
achieved by the isenthalpic expansion. Thus the vortex tube per-
forms the same function as the Joule-Thomson valve but pro-
duces a lower outlet gas temperature for a portion of the stream.
This apparatus could have application where gas pressure drop
is available, dewpoint control is needed, and the warm and cool
gases are recombined after liquid removal. This technology is
very space efficient and can be attractive, especially for offshore
processing applications.
FIG. 16-6
Concept of the Twister Process
16-7
Membrane Conditioning
Rubbery heavy hydrocarbon selective membranes are an-
other recent process development for use in fuel as condition-
ing. These membranes are different from conventional cellu-
losic membranes used for CO
2
removal from natural gas as
described in Section 21. These rubbery membranes preferen-
tially permeate the heavy hydrocarbons through the membrane
over methane, which is counterintuitive. The reason for this
reverse permeation behavior is that the rubbery membranes
separate gases based on solubility differences as compared to
diffusion rate differences in the molecules.
This separation difference allows these membranes to be
used for a variety of applications in the natural gas processing
and treating areas. The main application of these rubbery mem-
branes has been in remote area fuel gas conditioning to decrease
the heating value of raw fuel gas. This process allows the raw
gas to meet required engine specifications, thereby allowing op-
erators to run compression driven by gas engines and move the
gas from the remote location to the central processing facilities.
One version of the membrane will also remove CO
2
/H
2
S as well
as the heavy hydrocarbons. One company, MTR has dozens of
units installed in various offshore, onshore remote locations,
shale gas production areas and early production facilities all
over the world
4
(See Fig. 16-8).
Other commercial applications of the rubbery membranes
are in nitrogen rejection (to reduce N
2
content in natural gas),
and for LPG/NGL recovery from gas streams. In the nitrogen
rejection application, N
2
is the least permeable component in
the rubbery membrane, so the membrane can be used to reduce
the N
2
concentration in the pipeline gas, but in this case the
methane rich gas will be in the low pressure permeate along
with the heavier hydrocarbons. Therefore, the pipeline gas in
this case would have a high heating value recovery, while re-
jecting the nitrogen in the residue gas.
In LPG recovery, the fact that the heavy hydrocarbons per-
meate the membrane, can be effectively used to increase the
NGL content in the condensing loop, thereby allowing NGL re-
covery at higher condensing temperature. NGL recovery mem-
brane processes are most suitable for flare gas applications in
which NGL recovery is combined with power generation from
the clean residue gas produced by the rubbery membrane.
FIG. 16-8
Commercial Conditions for Rubbery Membranes
Type of Rubbery
Membrane
Pressure Max &
(Range) kPa (ga)
Flow rate
Max & (Range)
MSm
3
/day
Standard
Hydrocarbon selective
(SR Membranes)
8270
(345 8270)
3.4 (0.0028 3.4)
Polar Hydrocarbon
selective membranes
(Polaris Membranes)
6900
(345 6900)
0.28 (0.056 0.28)
Rotary Valve Fast Cycle PSA for
Fuel Gas Conditioning
A fast cycle Pressure Swing Adsorption (PSA) developed by
Xebec has been used successfully to condition gas.
5
This process
selectively adsorbs different gases by adsorbents when subject-
ed to varying pressures. Inlet gas flows up through a vessel
filled with a set of adsorbents at relatively high pressure. The
heavier hydrocarbons (C2+), inerts (CO
2
& N
2
) and water are
preferentially retained, while methane is not. Before the bed
becomes saturated, it is switched out of adsorbing mode, and
the pressure is reduced. The lower pressure releases the impu-
rities from the adsorbent as exhaust gases. The adsorbent bed
is then re-pressurized and is ready to repeat the cycle.
To ensure a continuous supply of product gas, the process
uses multiple adsorbent beds, which are cycled in a sequential
manner through identical operating cycles. The system con-
tains slowly rotating, multi-port, selector valves, which are cen-
tral to the operation of the unit. These valves are used to create
a very efficient, rapid PSA cycle, which results in smaller equip-
ment packages with reduced capital and installation costs than
a traditional PSA would require.
FIG. 16-7
Basic Design of a Vortex-Tube Device
16-8
STRAIGHT REFRIGERATION
NGL RECOVERY
The straight refrigeration process is quite flexible in its ap-
plication to NGL recovery. As outlined in the previous section,
the process can simply be used for dew point control when mod-
est liquid recovery is needed or desired. Alternatively, the pro-
cess can be used for high propane recovery and, in the case of
rich gases, for reasonable quantities of ethane recovery. The
recovery level is a strong function of the feed gas pressure, gas
composition and temperature level in the refrigeration chiller.
Fig. 16-9 shows curves for estimating the recovery achievable
as a function of temperature and gas richness for a given pro-
cessing pressure. (Liquid content in this figure is propane plus.)
Generally speaking, higher recovery efficiencies can be achieved
with richer feed gas. The straight refrigeration process is typi-
cally used with a glycol injection system. This configuration is
limited in the temperature of operation by the viscosity of the
glycol at the lower temperatures. Also, refrigeration is typically
provided by propane refrigeration which is limited to 42C re-
frigerant at atmospheric pressure and thus a processing tem-
perature of about 40C. In order to go lower in processing tem-
perature, upstream dehydration and alternative refrigeration
systems must be considered.
Fig. 16-10 illustrates the ethane recovery efficiency which
can be expected. As with propane recovery, for a given tempera-
ture level, higher extraction efficiency can be achieved with
richer gas. However, ethane recovery of over 30% can be
achieved from a gas as lean as 0.4 m
3
/1000 m
3
gas (C
3
+). Fig.
16-11 illustrates the effect of gas pressure on plant performance
in propane plus recovery operation.
Refrigeration Process Alternatives
There are many variations in the straight refrigeration pro-
cess. Fig. 16-12 illustrates four of the most common variations.
In the first scheme the gas is cooled against the residue gas and
FIG. 16-10
Recovery Efficiency, Ethane Plus
5
FIG. 16-11
Effect of Gas Conditions on Propane Recovery
5
FIG. 16-9
Recovery Efficiency, Propane Plus
5
16-9
FIG. 16-12
Refrigeration Process Alternatives
6
16-10
the cold separator liquid before being chilled with refrigeration.
This scheme uses a top-feed fractionator with the overhead be-
ing recompressed and recycled to the inlet. The use of the liquid
/feed gas exchanger helps reduce the chiller load. In this case,
the residue gas from the cold separator has a dew point of the
cold separator operating temperature.
The second scheme also uses a top-feed fractionator, but the
cold separator liquid is fed directly to the fractionator. This
fractionator operates with a lower overhead temperature which
justifies exchange with the refrigeration system. The overhead
after being warmed is recompressed and blended with the resi-
due gas from the cold separator. In this configuration the frac-
tionator overhead usually raises the residue gas dew point
somewhat. The cold separator temperature must be set to en-
sure that the desired dew point specification of the combined
stream is achieved.
The third process uses a refluxed fractionator. This type of
design usually has the highest liquid recovery efficiency, but
has a higher cost due to the overhead reflux system. The fourth
variation can be used where the cold separator liquid can be
pumped and the stabilizer run at an elevated pressure. This can
eliminate the need for a recompressor.
Any one or a combination of the following conditions:
Higher separator pressure
Richer gas
Recovery limited to propane-plus
will lead to higher recycle/recompression rates. This results in
more refrigeration power, more recompressor power, more frac-
tionator heat, and larger equipment. These conditions favor the
second and third schemes of Fig. 16-12.
Any one or a combination of the following conditions:
Lower separator pressure [around 4100 kPa (ga)]
Leaner gas (below 0.4 m
3
/1000 m
3
gas C3+)
Recovery includes ethane
will lead to lower recycle/recompression rates. These conditions
favor the first scheme in Fig. 16-12, or the fourth scheme if the
separator pressure is not higher than 2750-3100 kPa (ga). Sep-
arator pressure below 2750 kPa (ga), expecially with lean gas,
will result in poor product recovery.
Regardless of the exact configuration employed, the capacity
of the specific refrigeration system varies directly with refriger-
ant condensing temperature and evaporating temperature.
Condensing temperature is set by the condensing medium
available at the plant site, and the process chiller temperature
is set by the refrigerant evaporating temperature. Refrigerant
power requirements vary with condensing and evaporating
temperatures. Lower condenser temperature and higher evapo-
rating temperature require lower power per unit of refrigera-
tion required. For a given refrigeration load, power and con-
denser duties can be found in Section 14 for a variety of
refrigerants.
IPOR
SM
Process
The Lummus Technology/Randall Gas Technologies licenses
the IPOR
SM
process, a refrigeration based process, recovers
99.8+% C
3
from gas streams with essentially no C
2
recovery.
FIG. 16-13
IPOR
SM
Process
16-11
This design incorporates a closed loop propane refrigeration
cycle coupled with an open loop ethane mixed refrigeration cy-
cle. This enables lower temperature operation and provides a
reflux stream to the fractionation column, the combination of
which produces high recovery for relatively low power consump-
tion.
The process, shown in Fig. 16-13, can operate over a wide
range of pressures, depending on the feed conditions and re-
quired recoveries, and is well suited to handling rich feed gas.
The optimum feed pressure is 1725 to 3790 kPa (ga). The pro-
cess has an outlet pressure about 170 kPa lower than the inlet
pressure, so recompression will be required to meet higher resi-
due gas pressure battery limits.
The use of ethane-based refrigeration has been demonstrat-
ed to provide efficient high liquids recovery, particularly when
combined with simultaneous heat and mass transfer equip-
ment.
LEAN OIL ABSORPTION
NGL RECOVERY
Absorption is the physical process where a vapor molecule of
a lighter hydrocarbon component will go into solution with a
heavier hydrocarbon liquid (nonane, decane and heavier) and
be separated from the gas stream. The process can be operated
at ambient temperatures if only the heavier NGL products are
desired. A refrigerated system enhances the recovery of lighter
hydrocarbon products such as ethane and propane; lighter hy-
drocarbon products such as ethane and propane; however,
ethane recovery is limited to less than 50% to avoid having to
remove excessive amounts of methane in the Rich Oil Demetha-
nizer, which would lead to excessive solvent circulation. The
absorbing fluid (lean oil) is usually a mixture of paraffinic com-
pounds having a molecular mass between 100 and 200 assum-
ing limited ethane recovery.
Lean oil absorption processes have the advantage that the
absorber can operate at essentially feed gas pressure with min-
imal loss of pressure in the gas stream which exits the process.
Plants, whether ambient or refrigerated, are constructed of car-
bon steel. This type of process was used from the early part of
the 20th century and plants are still in use today. However,
most lean oil plants have been shut down or replaced with more
modern straight refrigeration or turboexpander process plants.
The lean oil process requires large processing equipment with
excessive energy requirements. Lean oil absorption units are
still used in many refinery operations.
Process Considerations
The desired composition of the lean oil is determined by the
absorber pressure and temperature. The optimum molecular
mass lean oil is the lowest mass oil which can be retained in the
absorber with acceptable equilibrium losses to the residue gas.
Lean oil absorption plants operating without refrigeration will
require a higher molecular mass oil, usually in the 150-200 mo-
lecular mass range. Refrigerated lean oil absorption systems
can operate with an absorbing medium as low as 100 molecular
mass with proper design.
Since the absorption is on a molar basis, it is desired to con-
tact the gas stream with the maximum number of moles of lean
oil to maximize the recovery of products from the gas. However,
the circulation rate is in units of volume, cubic meters per hour.
Therefore, a plant designed to circulate a heavier molecular
mass oil can circulate more moles of oil with the same equip-
ment if the molecular mass is lowered.
Many absorption oil recovery plants designed to originally
operate at ambient temperatures have been modified to include
a refrigeration system that allows both the lean oil and the gas
to be chilled before entering the absorber. The reduced temper-
ature increases the absorption and allows circulation of less oil
of lower molecular mass because the vaporization rate into the
residue gas is reduced. Oil is also lost with the NGL product. Oil
losses with the product can be minimized by improving frac-
tionation in the lean oil still. Many refrigerated lean oil absorp-
tion plants can recover enough heavy ends from the gas stream
to offset oil losses from the absorber, thereby making its own
absorption oil.
If the gas stream contains compounds that cause the ab-
sorption oil molecular mass to exceed design, a lean oil stripper
can be used on a side stream of circulating lean oil to remove
the heavy components. It is important to maintain the molecu-
lar mass of the absorption oil at the design value because the
circulating equipment, heat exchangers, and distillation pro-
cess are designed to utilize a particular molecular mass fluid.
Refrigerated Lean Oil
Fig. 16-14 shows a typical refrigerated lean oil absorption
process. The actual equipment configuration changes with dif-
ferent gas feeds and product recoveries.
Raw gas enters the plant inlet separator upstream of the
main process where inlet liquids are separated. The gas then
enters a series of heat exchangers where cold process gas and
the refrigerant reduce the feed gas temperature. This reduction
in temperature results in condensation of the heavier hydrocar-
bons in the inlet gas.
The gas is then fed to the bottom of the absorber where it
flows upward countercurrent to the lean oil which is introduced
at the top of the column. The lean oil has also been chilled to aid
in NGL absorption. This column has trays or packing which
increase the contact of the gas and lean oil. The lean oil physi-
cally absorbs the heavier hydrocarbons from the gas. The light-
er components stay in the gas and leave the top of the absorber.
The oil and absorbed hydrocarbons leave the bottom of the ab-
sorber as rich oil.
FIG. 16-14
Refrigerated Lean Oil Absorption
8
16-12
The rich oil flows to the Rich Oil Demethanizer (ROD) where
heat is applied to the rich oil stream to drive out the lighter
hydrocarbons which were absorbed. Some of the cold lean oil is
also fed to the top of the ROD to prevent loss of desirable NGLs
from the rich oil.
The rich oil from the ROD is then fed to a fractionation tow-
er or still. The still is operated at a low pressure and the NGLs
are released from the rich oil by the combination of pressure
reduction and heat addition in the still. The operation of the
still is critical to the overall plant operation as this is not only
the point where the desired product is produced, but the lean oil
quality from the bottom of the column is important in the ab-
sorption of NGLs in the absorber. The refrigeration required for
the oil and gas chilling and the heat inputs to the ROD and still
are the key parameters which must be controlled to operate a
lean oil plant efficiently.
LOW TEMPERATURE
NGL RECOVERY PROCESSES
Dew point control and mechanical refrigeration systems are
intended for applications where enough liquefiable hydrocar-
bons must be removed from the gas stream such that the gas
can be transported without liquids formation and/or where
some of the heavier components must be removed in order to
meet a maximum heat content specification. In these cases, the
goal is to meet the natural gas stream sales specification, not
recovery of liquids. In cases where the liquefiable hydrocarbons
present in a gas stream are more valuable as a liquid product
than as the heating value they contribute to the gas, more effec-
tive methods of liquids recovery are desired. The lean oil pro-
cesses described in the last subsection were the original tech-
nique used for this objective, but as already noted, this
technology has been superseded by more cost effective cryogen-
ic expansion process methods, which are covered in the follow-
ing sub-sections.
In order to condense and separate more ethane and propane
and thus achieve higher NGL recovery levels than are possible
with dew point control or propane-based refrigeration alone,
more gas cooling must be provided, which means cryogenic tem-
peratures (below 45C) are required. In order to achieve these
temperatures, a combination of pressure expansion and chilling
is used. The two types of cryogenic expansion processes in use
today are:
1. Adiabatic J-T expansion across a control valve
2. Isentropic expansion through a turboexpander
3. Variants of each system above with supplemental me-
chanical refrigeration
The turboexpander processes are the most widely used;
however, J-T expansion type processes are sometimes used for
lower recovery levels in unattended or remote locations, and
usually for gas flow rates less than 0.566 MSm
3
/day. Turboex-
pander type processes are used for higher ethane and propane
recovery levels with larger gas inlet flow rates where the cost of
the additional equipment is easily justified by the value of the
additional liquids recovered.
One limitation on the recovery of ethane and heavier prod-
ucts is the effect of liquids extraction on the heat content of the
residue gas. The presence of inerts in the feed gas may limit the
amount of ethane that can be removed and still meet the mini-
mum heating value specification for the residue gas.
Another consideration is the liquids recovery level that is
justifiable for a given gas composition and plant location. If eth-
ane can be recovered at a good price differential over its value
as a fuel in the residue gas, then the additional compression
and complexity of a high ethane recovery facility may be justi-
fied. If there is no local market for ethane, or no way to trans-
port high vapor pressure product to market, then only propane
and heavier components may have an economic incentive for
recovery. The plant may be designed for initial operation in a
propane recovery mode, with future conversion to ethane recov-
ery if or when the local demand for ethane justifies the expense
of additional compression. Many plants in North America are
designed for dual mode operation in which the operator can se-
lect high ethane recovery or ethane rejection, while still main-
taining high propane recovery. These plants can be switched,
during normal operation, as often as necessary between ethane
recovery and ethane rejection, in response to changes in the
processing margin. Historically, recovery of propane and heavi-
er components is almost always justified.
Propane Plus NGL Recovery Compared
to Ethane Plus NGL Recovery
Ethane is more volatile than the propane and heavier com-
ponents in a natural gas stream so generally deeper cooling is
required to condense the ethane fraction than is needed to con-
dense only the propane and heavier components. The tempera-
tures needed for ethane recovery are thus colder than those
needed for propane recovery regardless of the process design.
This means that more compression power is required for high
ethane recovery than for propane recovery only because of the
higher pressure differential required to produce lower tempera-
tures across the cryogenic sections and/or refrigeration system.
The colder operating temperatures for ethane recovery
mean that more of the plant will be constructed of stainless
steel or aluminum than may be needed for a facility which only
recovers propane. Another consequence of the colder tempera-
tures is that the allowable concentration of CO
2
in the feed to
avoid CO
2
freezing will be less for an ethane recovery design
than for a propane recovery design.
As stated previously, many process designs can operate in
either an ethane recovery mode or in an ethane rejection (pro-
pane recovery) mode, as long as the equipment design tempera-
tures are acceptable for both modes and the compression power
is sufficient to support the desired ethane recovery level. An
ethane recovery design will not be as efficient for propane re-
covery compared to a propane recovery specific design. It is,
however, possible to combine a very high efficiency ethane re-
covery design with a very high efficiency propane recovery de-
sign, if this need is identified during the design phase.
9
There is always an incentive to minimize the power, emis-
sions, and capital cost for the facility for a given recovery level
of ethane and/or propane. A simple J-T expansion type process
may have very good economic performance for a given loca-
tion, flow rate and product price, but it will have low process
recovery efficiency. This technology is therefore normally re-
stricted to relatively small plants. Typical J-T cryogenic pro-
cess configurations are described in the next sub-section. A
good understanding of J-T operation will help in understand-
ing the expander plant designs in the subsequent sub-sec-
tions. Turboexpander plants are operated in a J-T mode before
the turboexpander is started up and when the turboexpander
is unavailable. Examples are provided later showing the dif-
ference in product recovery for operation in J-T mode com-
pared to turboexpander mode.
16-13
FIG. 16-16
J-T Process with Mechanical Refrigeration Recovery
INLE T GAS
NGL P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
DE ME THANIZE R
J -T VALVE
COLD
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
ME CHANICAL
R E FR IGE R ATION
FIG. 16-15
J-T Process for Propane Recovery
INLET GAS
LPG PRODUCT
GAS
RESIDUE
COMPRESSOR
DEETHANIZER
J-T VALVE
COLD
SEPARATOR
HEAT EXCHANGE
RESIDUE
16-14

INLE T GAS
FR OM DE HY
LP G P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
DE E THANIZE R
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
FIG. 16-17
Simple Turboexpander Process for Propane Recovery
DE E THANIZE R
INLE T GAS
FR OM DE HY
LP G P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
S UBCOOLE R
R E S IDUE
FIG. 16-18
GSP Process for Propane Recovery
16-15
J-T EXPANSION
The general concept for the Joule-Thomson (J-T) Expansion
design is to chill the gas by expanding the gas across a J-T
valve. With appropriate heat exchange and large pressure dif-
ferential across the J-T valve, cryogenic temperatures can be
achieved resulting in acceptable extraction efficiencies. The
main difference between the J-T design and turboexpander is
that the gas expansion is adiabatic across the valve and is near-
ly isentropic across the turboexpander. The outlet temperature
is colder for a given pressure ratio for the turboexpander than
for a J-T valve. The result is that the J-T design will have much
lower ethane or propane recovery for a given amount of residue
compression power than the turboexpander process, but the
process is very simple, which helps make it suitable for some
applications.
Specifically, the J-T process is used for situations where the
higher efficiency due to the lower temperatures generated and
the work recovered via the turboexpander driven compressor do
not offset the increased cost & complexity involved, such as:
1. The feed gas rate is low and/or,
2. There is a relatively low ethane and propane recovery re-
quirement
3. The unit will be located in a remote or unattended loca-
tion
4. Where a broad range of inlet gas flow rates and composi-
tions can be expected.
Typical Process Flow for J-T Process
Fig. 16-15 illustrates a typical process flow arrangement for
a J-T expansion process. In order to effectively use the J-T pro-
cess, the gas must be at a high inlet pressure. Pressures over
6900 kPa (abs) are typical in these facilities. If the gas pressure
is too low, inlet compression is necessary or sufficient expansion
chilling will not be attained. The gas must first be dried to en-
sure that no water enters the cold portion of the process. Typi-
cally, molecular sieves or alumina are used for the drying.
Methanol injection has been used in a few plants successfully
but can be an operating problem because the dew point is
too close to normal operating temperatures.
After drying, the gas is cooled by heat exchange with the
cold residue gas and also by heat exchange with the demetha-
nizer mid-tower liquids and in some cases the liquids from the
cold separator. After chilling, the gas is expanded across the J-T
valve and sent to the cold separator . The liquid from this sepa-
rator is the feed to the demethanizer. This demethanizer col-
umn is needed because the separator liquids contain too much
methane and ethane to meet a liquid product specification for
downstream fractionation. Usually this tower is a cold, top feed
design. The cold liquid is demethanized to the proper specifica-
tion in this tower. The cold overhead product from the demetha-
nizer is used to cool the feed and is then recompressed as neces-
sary for residue sales. When operated in an ethane recovery
mode, enough heat is provided to the bottom reboiler to limit
the methane content to an acceptable level. When operated in
an ethane rejection mode, the bottom reboiler heat input is ad-
justed to achieve an acceptable ethane content specification and
the ethane is sent overhead and on to the residue gas sales
point.
The liquids recovery for this process is determined by the
pressure ratio across the J-T valve and the quantity of heat ex-
change surface included in the plant heat exchangers. The pro-
cess can operate over a wide range of feed gas conditions and
produce specification product. The process is thus very simple
to operate and is often operated as an unattended or partially
attended facility. The pressure ratio across the J-T valve is
provided by the inlet gas pressure available upstream of the J-T
valve and the column pressure downstream of the J-T valve.
The power requirement for the residue gas compressor will in-
crease as the column pressure is lowered. Higher relative resi-
due compression power level results in higher pressure ratio
across the valve and more cooling, so the recovery of the more
volatile components increases. When the pressure ratio reaches
3:1, additional cooling can usually be supplied more efficiently
using a refrigeration loop as described in the following sub-sec-
tion. (Each megawatt of refrigeration will provide more incre-
mental recovery than one megawatt increase in residue com-
pressor power.)
Refrigerated J-T Expansion Process
In some cases the feed gas is not at high enough pressure or
the gas is rich in liquefiable hydrocarbons. Mechanical refrig-
eration can be added to the J-T process to increase recovery of
ethane and propane without adding a turboexpander. Fig. 16-
16 shows a J-T process where a small packaged refrigeration
system has been added to chill the feed gas. Another process
variation is shown in this figure. The gas in this design is ex-
panded downstream of the cold separator, taking advantage of
the cooling provided by the refrigeration system and feeding the
rich liquids separated at high pressure lower in the column.
(The optimum location for the J-T valve is dependent on the
inlet gas pressure and composition.) The advantage of exter-
nal refrigeration is that lower feed pressure can be used or,
alternatively, the demethanizer can be operated at a higher
pressure thus reducing residue compression. For rich gas
streams, using external refrigeration is more effective than in-
creasing inlet or residue compression. The total compression
power requirement (refrigeration plus residue plus inlet) may
be minimized with the use of a refrigeration system although
complexity is increased. The optimum combination of compres-
sion power versus recovery level must be determined by process
simulation of multiple equipment arrangements.
The J-T process, whether refrigerated or non-refrigerated,
offers a simple, flexible process for relatively low ethane and
propane recovery levels where a turboexpander is not justified.
However, modern turboexpanders have proven to be so benefi-
cial and reliable that almost all cryogenic plants are designed
around turboexpander processes.
TURBOEXPANDER PROCESSES
The modern turboexpander based cryogenic gas plant was
commercialized in the mid-1960s. Since then almost all new
high pressure NGL recovery plants built have employed a vari-
ant of this basic process. Turboexpander process designs use
the feed gas pressure to produce needed refrigeration by expan-
sion across a turbine (turboexpander). The turboexpander re-
covers useful work from this gas expansion. Typically the ex-
pander is linked to a centrifugal compressor to recompress the
residue gas from the process. Because the expansion is near is-
entropic, the turboexpander lowers the gas temperature signifi-
cantly more than expansion across a J-T valve. Details of the
turboexpander equipment are in Section 13.
This extraction of energy from the inlet gas using a turboex-
pander and its application to the booster compressor is why the
inlet gas stream can be cooled more for a given residue compres-
16-16
sor power level than in a J-T process. More cooling means more
condensation of the desired products so higher product recovery
can be achieved for a given residue compressor size and power
level.
PROPANE RECOVERY PROCESSES
The simplest (and least efficient) turboexpander process for
propane plus recovery uses a top feed, non-refluxed deethanizer
column as shown in Fig. 16-17. Inlet gas is cooled then sepa-
rated in to vapor and liquid phases at the cold separator. The
vapor is routed to the expander inlet and the liquids are routed
on level control through the inlet gas/liquids exchanger to a
mid-column feed. The expander outlet is fed to the top of the
deethanizer column where the vapor and liquids are separated.
The vapor exits the column and the condensed liquids from the
expander outlet are fractionated to separate the propane and
heavier components desired in the bottoms liquids from the
methane and ethane that were condensed as the vapor was
cooled in the expander. Note that any components not con-
densed in the expander outlet are lost to the residue gas stream.
For a propane recovery mode plant, maximum practical recov-
ery is limited to around 90% due to these losses at the top of the
column.
The cold separator liquids also contain methane and ethane
as well as propane and heavier components so the liquids are
routed to a mid-column feed for fractionation as well. The col-
umn bottoms are heated in a reboiler to a temperature suffi-
cient to boil off the methane and ethane with minimum loss of
propane to the overhead stream. The required bottoms temper-
ature is always warmer than the plant inlet gas temperature so
an external source of heat is needed for the reboiler. Hot oil,
steam, or hot compressor discharge gas are commonly used for
this service.
As higher ethane and propane recovery levels have been de-
sired, more efficient NGL recovery designs have been devel-
oped. The focus of these designs is to produce more reflux with
leaner composition for the fractionation column so that the com-
ponents not condensed at the expander outlet feed to the col-
umn can still be recovered while reducing the compression
power requirement.
GSP Design for Propane Recovery
The configuration for a Gas Subcooled Process (GSP) design
as originally developed by Ortloff is shown in Fig. 16-18 below.
The front end heat exchange arrangement is identical to the
simple turboexpander process but the deethanizer column is ex-
tended and stages are added above the expander feed. Reflux for
this new column section is provided by taking a portion of the
expander feed separator vapor and condensing and subcooling it
at inlet pressure prior to let down on flow control to the column
top feed. The reflux stream is very effective in condensing much
of the propane that was not condensed in the expander outlet.
The GSP design in propane recovery (ethane rejection) mode pro-
vided a much needed improvement in efficiency over the simple
turboexpander design,
10
but still has lower efficiency than de-
signs developed specifically for propane recovery.
Improved Open Art Propane
Recovery Processes
The simple turboexpander process and GSP design described
above are not as efficient for propane recovery as any of the proc-
ess designs developed specifically for high propane recovery
(ethane rejection). The simplest high propane recovery process is
an overhead recycle (OHR) design shown in Fig. 16-19. This
process configuration uses an absorber column and a separate
deethanizer fractionation column to achieve the desired recov-
ery. The overhead from the deethanizer is condensed and used to
absorb propane from the expander outlet stream. The expander
outlet stream is fed to the bottom of the absorber column. As the
reflux stream is warmed and partially vaporized by the expander
outlet vapor, the vapor is cooled and additional propane is con-
densed. The absorber overhead stream is colder than either of
the feeds due to the absorption effect. The maximum propane
recovery is limited by vaporization of propane from the reflux
stream. This configuration provided much more efficient recov-
ery of propane than the original simple turboexpander design.
Example Comparison for
Propane Recovery Designs
The hypothetical example shown in Fig. 16-20 is presented
to demonstrate the differences in the NGL recovery designs for
propane plus recovery at constant residue compression power.
The results for recovery of propane and heavier components us-
ing J-T expansion, simple turboexpander, GSP, and overhead
recycle processes are presented. Also tabulated are typical col-
umn pressures and temperatures. The purpose here is to dem-
onstrate the differences in operating conditions and power re-
quirements for different processes.
ETHANE RECOVERY PROCESSES
The simple turboexpander process was originally used for
ethane recovery. When used for ethane recovery, the column
pressure is lowered to provide a higher expansion ratio across
the expander, thus generating more cooling and condensing
more of the ethane in the expander outlet stream but with in-
creased residue gas compression. The amount of heat needed in
the bottom of the column is substantially less than for the eth-
ane rejection mode because only the methane must be vapor-
ized from the column bottoms liquids. The resulting column
bottoms temperature is usually cool enough so that inlet gas
can be used to supply some if not all of the required column
duty. A column side heater may also be used to provide more
inlet gas cooling while distributing the total heat input more
uniformly to the column.
The simple turboexpander process for ethane recovery is
shown in Fig. 16-21. Note the similarity to the J-T process with
the J-T valve replaced by the expander. Dry feed gas is first
cooled against the residue gas and used for side heating of the
demethanizer. Additionally, with richer gas feeds, mechanical
refrigeration is often needed to supplement the gas chilling.
The chilled gas is sent to the cold separator where the con-
densed liquid is separated, flashed and fed to the mid-point feed
of the demethanizer. The vapor flows through the turbo ex-
pander and feeds the top of the column. A J-T valve is always
installed in parallel with the expander. This valve can be used
for the plant inlet gas flow if the expander is out of service.
When the expander is out of service, the plant operates like the
J-T plant described previously, so the ethane recovery decreas-
es substantially.
In this configuration the maximum ethane recovery is lim-
ited to about 80%. The cold separator must be operated at a low
temperature to maximize recovery. Often the high pressure and
low temperature conditions are near the critical point of the gas
making operation unstable when very small changes in separa-
tor temperature result in large changes in the vapor volume to
the expander. Another problem with this design is if there is
more than a few tenths of a percent CO
2
in the feed. The CO
2

16-17
DE E THANIZE R
INLE T GAS
FR OM DE HY
LP G P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
CONDE NS E R
R E FLUX P UMP S
ABS OR BE R
FIG. 16-19
Overhead Recycle Process for Propane Recovery
A B C D
Process Design J-T SIMPLE GSP OHR
Calculated Ethane Recovery % 0.20% 0.67% 0.75% 0.82%
Calculated Propane Recovery % 22.8% 77.1% 86.4% 94.8%
Calculated Butanes Recovery % 55.5% 97.5% 98.4% 99.8%
Total Liquids Recovered m
3
/day 1546 2146 2191 2266
Expander Power kW 0 2386 3145 3667
Column Overhead Temperature C 28 64 69 75
Column Bottoms Temperature C 160 116 109 107
Column Reboiler Duty MW 3.0 4.5 5.1 5.6
Residue Gas Flow Rate MSm
3
/day 6.88 6.71 6.68 6.68
Notes:
1. C2/C3 Ratio set at 2.0 mol% all designs
2. Inlet Gas Flow 7.1 MSm
3
/day, 6.6% Ethane, 2.8% Propane, 43C, 6900 kPa (ga)
3. Residue Delivery 49C, 6900 kPa (ga)
4. Residue Compressor Power set at 8799 kW
FIG. 16-20
Process Design Comparison for Propane Recovery
16-18
FIG. 16-21
Simple Turboexpander for Ethane Recovery

INLE T GAS
NGL P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
DE ME THANIZE R
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
FIG. 16-22
GSP Process for Ethane Recovery
DE ME THANIZE R
INLE T GAS
FR OM DE HY
NGL P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
S UBCOOLE R
R E S IDUE
16-19
can solidify at operating temperatures found in the top stages of
the demethanizer. Later in this Section under Solid Forma-
tion there is more information on the CO
2
freezing problem
and the methods for prediction of solid CO
2
formation.
GSP Design for Ethane Recovery
The Gas Subcooled Process (GSP) was developed to over-
come the problems encountered with the conventional expander
process for ethane recovery. This process, shown in Fig. 16-22,
deviates from the original expander process in several ways. A
portion of the gas from the cold separator is sent to a heat ex-
changer where it is totally condensed and subcooled at inlet gas
pressure using the cold column overhead stream. This stream is
then flashed across a flow control valve to the top feed of the
demethanizer providing reflux to the demethanizer. Because
the stream is subcooled, only a small fraction of the stream
flashes at the reflux feed point.
The expander feed is sent to the tower several stages below
the top of the column. Vapor rising from the expander feed will
contain a significant amount of ethane, which is condensed by
the colder reflux stream.
In this process the column overhead is warmed up and the
column pressure is increased significantly without sacrificing
liquid recovery, due to the use of the subcooled reflux and the
cold separator now operates at a much warmer temperature,
well away from the system critical. The flow rate through the
expander is less than in a non-refluxed design, but the vapor is
much warmer, so the expander power is actually higher for the
GSP design than for a non-refluxed design. The residue com-
pression power is always much less at a given ethane recovery
level than for the non-refluxed design so the non-refluxed de-
sign is no longer used.
Example Comparison of Ethane
Recovery Designs
The hypothetical example presented earlier for propane re-
covery is presented in Fig. 16-23, but for ethane recovery to
demonstrate the differences in the ethane recovery for several
ethane recovery designs. The design basis assumptions from
the propane plus recovery comparison are used here with the
additional constraint of 2.0% max methane/ethane ratio for the
bottoms product. For ethane recovery, only the J-T, simple tur-
boexpander, and GSP options are tabulated since the overhead
recycle process is a propane recovery design.
LICENSED NGL
RECOVERY PROCESSES
Most ethane recovery plants today are screened initially us-
ing a GSP design to determine rough power requirements and
to see if refrigeration is needed. A GSP design may provide good
results when the CO
2
in the feed is low, the gas is not too rich,
and only the ethane recovery mode of operation is desired. How-
ever, there are many design basis requirements for which a li-
censed design is advantageous. Many locations have require-
ments that cannot be met with the process designs described
above. Some of these needs can be met using licensed designs
available from several sources.
FIG. 16-23
Process Design Comparison for Ethane Recovery
A B C
Process Design J-T SIMPLE GSP
Calculated Ethane Recovery % 8.2% 59.3% 84.0%
Calculated Propane Recovery % 27.1% 93.8% 98.7%
Calculated Butanes Recovery % 56.5% 99.2% 99.8%
Total Liquids Recovered m
3
/day 1136 2753 3203
Expander Power kW 0 3096 2820
Column Overhead Temperature C 29 83 94
Column Bottoms Temperature C 113 34 29
Column Reboiler Duty MW 2.1 2.1 1.8
Residue Gas Flow Rate MSm
3
/day 6.82 6.37 6.26
Notes:
1. C2/C3 Ratio set at 2.0 mol% all designs
2. Inlet Gas Flow 7.1 MSm
3
/day, 6.6% Ethane, 2.8% Propane, 43C, 6900 kPa (ga)
3. Residue Delivery 49C, 6900 kPa (ga)
4. Residue Compressor Power set at 9620 kW
16-20
FIG. 16-24
IOR

Propane Recovery Process


DE E THANIZE R
INLE T GAS
FR OM DE HY
LP G P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
E XP ANDE R /
COMP R E S S OR
E XP ANDE R
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
CONDE NS E R
R E FLUX P UMP S
ABS OR BE R
A
DEETHANIZER
INLET GAS
FROM DEHY
LPG PRODUCT
GAS
RESIDUE
COMPRESSOR
EXPANDER /
COMPRESSOR
EXPANDER
SEPARATOR
HEAT EXCHANGE
RESIDUE
CONDENSER
REFLUX PUMPS
SEPARATOR
FIG. 16-25
SCORE

Propane Recovery Process


16-21
IOR

and SCORE

Propane
Recovery Processes
The basic overhead recycle process has been modified by
Ortloff to make better use of the refrigeration available in the
feed streams. This improved version known as IOR

process as
shown in Fig. 16-24 makes a few improvements to the basic
overhead recycle process. In this process the reflux for the
deethanizer is produced in an absorber overhead system, which
produces reflux for both towers. Adding this reflux to the deeth-
anizer column helps minimize the amount of propane in the
reflux thus providing a leaner reflux, so the column pressure
can be higher for a given propane recovery. The absorber bot-
toms are heated against the feed before being sent to the deeth-
anizer thus cooling the inlet gas.
This process is typically designed for 99% propane recovery,
while rejecting 98% of the ethane. The bottoms temperature is
set to limit ethane in the bottoms to a level which will allow the
propane product at the downstream fractionator to meet a va-
por pressure specification. The improved heat integration and
reflux utilization result in a 5-15% reduction in the residue
compressor power requirement over the basic overhead recycle
design at the same recovery level.
Ortloff later developed a simplified version of the IOR

proc-
ess by combining the absorber and deethanizer into a single
column with a vapor side draw below the expander feed to pro-
duce reflux. This process shown in Fig. 16-25 is known as the
Single Column Overhead Recycle (SCORE

) process.
IPSI Enhanced NGL Recovery Process
Another improvement of the turboexpander-based NGL
GSP design is the IPSI Enhanced NGL Recovery

Process
11

shown in Fig. 16-26. This process utilizes a slip stream from or
near the bottom of the distillation column (demethanizer) as a
mixed refrigerant. The mixed refrigerant is totally or partially
vaporized, providing refrigeration for inlet gas cooling other-
wise normally accomplished using an external refrigeration
system. The vapor generated from this self-refrigeration cycle
is specifically tailored to enhance separation efficiency, then is
recompressed and recycled back to the bottom of the tower
where it serves as a stripping gas. The innovation not only re-
duces or eliminates the need for inlet gas cooling via external
refrigeration, but also provides the following enhancements to
the demethanizer operation:
Lowers the tower temperature profile, permitting better
energy integration for inlet gas cooling via reboilers, re-
ducing heating & refrigeration requirements.
Reduces and/or eliminates the need for external reboiler
heat, thereby saving fuel plus refrigeration.
Enhances the relative volatility of the key components in
the tower when operated at a typical pressure, improving
separation efficiency and NGL recovery; or allowing in-
creased tower pressure with lower recovery efficiency,
reducing the residue gas compression requirements.
LPG-MAX
SM
Process
The Lummus Technology/Randall Gas Technologies LPG-
MAX
SM
process recovers 99+% propane from gas streams with
essentially no ethane recovery. A simplified version of the
scheme is available that has lower recoveries (93 to 95%) with
lower capital cost.
This process, shown in Fig. 16-27, uses an absorber at high
pressure to minimize recompression. The pressure of the resi-
due gas leaving the booster compressor (driven by the expander)
will depend on feed pressure, and recovery required. Process
conditions are tailored to maximize pressure at booster outlet.
RSV

High Ethane Recovery Process


The limitation to ethane recovery for the GSP design is the
ethane content in the reflux liquids. The GSP reflux feed has
the same composition as the cold separator vapor so it contains
some ethane. A portion of this ethane will flash at the top feed,
thus generally limiting the recovery to around 93%, regardless
of the residue compressor power used. A leaner reflux stream is
needed to achieve recoveries higher than 93%. By adding stages
above the GSP reflux feed, and routing a condensed and sub-
cooled stream, created by recycling a small portion of the resi-
due gas back through the gas/gas exchanger and subcooler at
residue gas delivery pressure, to provide a lean reflux stream,
up to 99% ethane recovery can be achieved.
This Recycle Split Vapor (Ortloff RSV

) process design
shown in Fig. 16-27 is an efficient process in very high ethane
recovery designs, because the recycle rate is minimized by
maintaining the GSP reflux stream at the second feed for bulk
ethane recovery, and then only using enough recycle reflux at
the top feed to capture the ethane from the equilibrium losses
at the GSP reflux feed point.
NGL-MAX
SM
High Ethane Recovery Process
The Lummus Technology/Randall Gas Technologies NGL-
MAX
SM
process recovers 99+% ethane from gas streams with
essentially 100% propane-plus recovery. This process, shown in
Fig. 16-29, uses semi-lean and lean reflux to achieve very high
ethane recovery or to increase gas throughput at lower recover-
ies.
The process can be designed to run in propane recovery
mode (dual operation, ethane recovery, and propane
recovery). Switching from ethane recovery mode to propane
recovery mode will ensure that there is almost complete
ethane rejection with 99% C
3
+ recovery.
Other Licensed NGL Recovery Processes
Other examples of licensed designs include the following:
1. CO
2
content in the feed too high for the liquid product
CO
2
/C
2
ratio without having to reboil the column to a
CO
2
specification and lose ethane recovery. (Ortloff
CDC)
2. CO
2
content in the feed high enough to freeze using GSP
design, but inlet treating is not justified. (Ortloff RSV)
3. Retrofit of an existing plant is limited by the existing in-
let gas system throughput. (Technip DEER)
4. Dual mode and intermediate ethane recovery levels are
desired without loss of propane recovery. (Ortloff SRP)
5. Dual column designs for high propane recovery with flex-
ible ethane recovery (Lummus Randall Super Hy-Pro
STC & TTC)
6. Integration of NGL Recovery and LNG production are
desired. (Ortloff, IPSI)
16-22
7. Integration of NGL Recovery and Nitrogen Rejection are
desired. (Costain/Chart) This type of design is discussed
in more detail in the Nitrogen Rejection subsection.
AVOIDING COMMON
OPERATING PROBLEMS
Most operating problems with cryogenic NGL recovery
plants can be avoided with attention to detail in the design and
attention to cleanliness during construction and maintenance.
The following guidelines apply to both open art and licensed
designs.
1. Mole sieve dust from the dehydrators will plug the brazed
aluminum heat exchangers (BAHE) used in most cryo-
genic plants. Very high quality adequately sized filters
must be installed upstream of the heat exchangers. They
must be installed in a two by 100% configuration without
a bypass. Inlet gas must never be allowed to bypass the
filters, and differential pressure must be monitored. The
importance of good filtration and good operating practice
cannot be overemphasized. Sudden plugging of the filters
with mole sieve dust indicates disintegration of the mole
sieve itself or failure of the bed supports and must be ad-
dressed immediately to prevent permanent plugging of
the exchangers.
2. The mole sieve dehydration system cooling gas tempera-
ture should be cool enough to get the cooled dehy bed
down to as close to the inlet gas temperature as possible.
If the regenerated bed is too warm when placed into ser-
vice, an inlet gas temperature spike will result as the bed
is further cooled by inlet gas. The inlet gas will warm up
and create a disturbance throughout the cryo plant. This
disturbance will occur every time the dehys switch.
3. During commissioning, construction debris will typically
migrate from the column packing or trays to the liquid
draws from the column. The debris will then flow to the
side heaters/reboilers and accumulate at the strainer.
Two parallel 100% strainers with block valves on the liq-
uid draw lines will allow cleaning of the strainers with-
out a shutdown.
4. The initial design must include low point drains to allow
blow down of all free water prior to cool down. Dryout
lines from the dehys to the bottom of the column and to
the reflux system are typically installed to allow a flow of
warm dry gas through the entire plant to remove mois-
ture. A plant recycle line from the residue compressor
discharge back to the dehy system will allow dryout of
the plant with minimum flaring. Dryout must be ad-
dressed during the design phase of the project.
5. Changes in reflux flow rates or bottoms temperature set-
points may take several hours to settle out in process de-
signs where there are recycle type reflux systems. Opera-
tors must allow sufficient time for the plant to stabilize
when making even small changes as feed composition or
flow rate changes.
6. Any decrease in pressure ratio across expander the ex-
pander will reduce recovery. Any increase in pressure
drop across strainers or exchangers will result in lower
expander inlet pressure or higher column pressure and
normal recovery will not be achieved until the normal
pressures at the expander are reestablished. Gradual
increases in pressure drop on the inlet gas side of the
heat exchangers may indicate mole sieve dust getting
through damaged filter elements. Gradual increases in
pressure drop across the cold exchangers may indicate
deteriorating dehydration system performance and a
FIG. 16-26
IPSI Enhanced NGL Recovery Process
16-23

DE ME THANIZE R
INLE T GAS
NGL P R ODUCT
GAS
R E S IDUE
COMP R E S S OR
E XP ANDE R /
COMP R E S S OR
COLD
S E P AR ATOR
HE AT E XCHANGE
R E S IDUE
S UBCOOLE R /
GAS-GAS
E XCHANGE R
FIG. 16-28
RSV

Ethane Recovery Process


FIG. 16-27
LPG-MAX
SM
Propane Recovery Process
16-24
FIG. 29
NGL-MAX
SM
Ethane Recovery Process
FIG. 16-30
Four-column Fractionation System
16-25
gradual accumulation of ice. Methanol injection is typi-
cally used to determine if the pressure drop is due to ice
or particulates. Note that if operating below 98C,
Methanol will freeze.
7. One treating requirement that becomes important to con-
sider in design when using these processes is to confirm
if mercury can be present in the feed gas, and if so to add
treating facilities to ensure any mercury present is re-
moved to very low levels. A typical specification will be
<0.1 microgram per cubic meter (<0.01 ppbv)) of inlet
gas. See Section 21 Mercury Removal for more infor-
mation).
8. When molecular sieves are regenerated species co-ad-
sorbed on the sieve are also removed. This means that
when the newly regenerated bed is brought back on line
there will be a short period where the species co-adsorbed
such as any acid gases still present and more important-
ly, light hydrocarbons like ethane, propane and butane,
are re-adsorbed, and these species are removed from the
gas stream. If this will have a negative impact on the
downstream NGL recovery process, then a gas contain-
ing these species needs to be used as part of the regenera-
tion process to pre-saturate the bed prior to it being
placed back in drying service.
FRACTIONATION CONSIDERATIONS
The cryogenic plant fractionation column produces the spec-
ification C2+ or C3+ stream as a bottom product. This mixed
product then needs to be separated into usable products in a
series of one or more fractionation columns. This fractionation
may take place at the cryogenic plant facility or at some frac-
tionation facility down the pipeline. However, many gas plants
have onsite fractionation of the recovered NGL stream so that
ethane, propane, and butane can be sold separately from that
site.
10
If the NGL stream is an ethane plus stream, the first step is
to separate the ethane from the propane and heavier compo-
nents in a deethanizer. The propane is then separated from the
butane and heavier components in a depropanizer. If further
processing is desired the butane may be separated in a debuta-
nizer and the butanes further separated in a butane splitter
column. The butane splitter is only used when a differential
value can be realized for the isobutane versus the mixed butane
stream. A schematic of a four column fractionator is shown in
Fig. 16-30. Section 19 in the Data Book covers the specifics of
fractionation systems for NGL streams. Some markets will re-
quire a slightly different fractionation system arrangement in
which the deethanizer bottoms liquid is routed to an LPG col-
umn where the C3 and C4 LPG components are separated from
FIG. 16-31
Nine-stage Cascade Liquefaction Process
12
16-26
the C5+ components. The C3+C4 stream may then be further
fractionated in to separate C3 and C4 components if the local
LPG spec cannot be met without reducing the amount of one
component in the LPG product and selling that component sep-
arately.
LIQUEFIED NATURAL GAS
PRODUCTION
The principal reason for liquefying natural gas is the 600-
fold reduction in the volume which occurs with the vapor-to-
liquid phase change. This volume reduction is important in the
transportation and storage of the gas. In the liquid state, the
gas can be transported in discrete quantities, can be economi-
cally stored in tanks for use as required, and can be transported
long distances not feasible with gas pipelines.
Because methane is the primary component of natural gas,
the production of Liquefied Natural Gas (LNG) involves the
chilling of the entire natural gas feed stream to cryogenic tem-
peratures sufficient to totally condense the gas stream. Com-
mon to all LNG liquefaction processes is the need to pretreat
the gas to remove components, such as CO
2
and water, which
will solidify in the liquefaction step. The liquefaction unit also
has to remove hydrocarbon components, such as benzene and
cyclohexane, which can solidify. Two types of LNG facilities
have been developed: 1) large base load units for continuous
LNG production to export markets, and 2) small peak shaving
plants for gas distribution systems. The large scale based load
units are typically designed with emphasis on process efficien-
cy. In addition to the process units involved in the liquefaction
step, base load LNG plants tend to be large complex facilities
which involve product storage, loading and complete stand-
alone utility systems.
Peak shaving facilities differ from base load units in several
aspects. Peak shaving plants are much smaller, operate only a
portion of the year, and are often located near the point of use
for the gas. The design emphasis is thus on capital cost minimi-
zation rather than thermodynamic efficiency. In order to pro-
duce the low temperature necessary for liquefaction, mechani-
cal refrigeration systems are utilized. Four types of liquefaction
processes can be used to accomplish this refrigeration:
1. Cascade Refrigeration Processes
2. Mixed Refrigerant Processes
3. Precooled Mixed Refrigerant Processes
4. Turboexpander Based Process (Peak Shaving)
Each of these processes has been used for liquefaction facili-
ties with the PreCooled process being the predominant technol-
ogy in base load units.
As of 2010 in the United States these processes are all being
used in peak shaving units with the following distribution.
Mixed Refrigerant: 41
Expander Based: 14
Cascade or Nitrogen: 5
Total: 60
Cascade Refrigeration
The first LNG liquefaction units utilized the cascade refrig-
eration process. In the classical cascade cycle three refrigera-
tion systems are employed: propane, ethylene and methane.
Two or three levels of evaporating pressures are used for each
of the refrigerants with multistage compressors. Thus the re-
frigerants are supplied at eight or nine discrete temperature
levels.Using these refrigeration levels, heat is removed from
the gas at successively lower temperatures. The low level heat
removed by the methane cycle is transferred to the ethylene
cycle, and the heat removed in the ethylene cycle is transferred
to the propane cycle. Final rejection of the heat from the pro-
pane system is accomplished with either water or air cooling.
Early facilities used a closed methane refrigeration loop.
More modern designs use an open methane loop such as shown
in Fig. 16-31 where the methane used for refrigerant is com-
bined with the feed gas and forms part of the LNG product. The
efficiency and cost of the process is dependent on the number of
refrigeration levels provided in each refrigeration system.
The refrigeration heat exchange units traditionally were
based on shell and tube exchangers or aluminum plate fin ex-
changers. Newer designs incorporate plate fin exchangers in a
vessel known as core-in-kettle designs. A critical design ele-
ment in these systems is the temperature approach which can
be reached in the heat exchangers.
Mixed Refrigerant Processes
After initial developments of cascade LNG plants, the mixed
refrigerant cycle was developed to simplify the refrigeration
system. This system uses a single mixed refrigerant composed
of nitrogen, methane, ethane, propane, butane and pentane.
The refrigerant is designed so that the refrigerant boiling curve
nearly matches the cooling curve of the gas being liquefied. The
closeness of the match of these two curves is a direct measure of
the efficiency of the process.
The process (Fig. 16-32) has two major components: the re-
frigeration system and the main exchanger cold box. The cold
box is a series of aluminum plate fin exchangers which provide
very close temperature approaches between the respective pro-
cess streams. The low pressure refrigerant is compressed and
condensed against air or water in a closed system. The refriger-
ant is not totally condensed before being sent to the cold box.
The high pressure vapor and liquid refrigerant streams are
combined and condensed in the main exchanger. The condensed
stream is flashed across a J-T valve and this low pressure re-
frigerant provides the refrigeration for both the feed gas and
the high pressure refrigerant.
Removal of pentane and heavier hydrocarbons from the feed
gas is accomplished by bringing the partially condensed gas out
of the cold box and separating the liquid at an intermediate
temperature. The liquid removed is then further processed to
produce a specification C
5
+ product. Light products from this
separation are returned to the liquefaction system.
Precooled Mixed Refrigerant Process
The propane precooled mixed refrigerant process (Fig. 16-33)
was developed from a combination of the cascade and mixed re-
frigerant processes. In this process, the initial cooling of the feed
gas is accomplished by using a multistage propane refrigeration
system. The gas is cooled with this system to around 40C at
which point the gas is processed in a scrub column to remove the
16-27
FIG. 16-32
Mixed Refrigerant Liquefaction Process
13
FIG. 16-33
Propane Precooled Mixed Refrigerant Process
14
16-28
heavy hydrocarbons. The gas is then condensed in a two step
mixed refrigerant process. The chilling of the gas is accomplished
in a single, large, spiral-wound heat exchanger. This exchanger
allows extremely close temperature approaches between the re-
frigerant and the gas to be achieved.
The mixed refrigerant in this process is a lighter mixture
composed of nitrogen, methane, ethane and propane with a mo-
lecular mass around 25. The mixed refrigerant after recompres-
sion is partially cooled with air or water and then further cooled
in the propane refrigeration system. The partially condensed
refrigerant from the propane chilling is separated and the high
pressure vapor and liquid streams sent separately to the main
exchanger. The liquid is flashed and provides the initial chill-
ing of the gas. The high pressure vapor is condensed in the
main exchanger and provides the low level, final liquefaction of
the gas. As in the other processes, the LNG leaves the exchang-
er subcooled and is flashed for fuel recovery and pumped to
storage.
Dual Mixed Refrigerant Cycle Processes
Building on the success of the propane pre-cooled mixed re-
frigerant cycle Shell and Linde have both successfully commer-
cialized a dual or double mixed refrigerant cycle process. In this
process the multi-stage propane refrigeration system used to
cool the inlet gas and partially condense the mixed refrigerant
is replaced with a second warm mixed refrigerant tailored to
provide this cooling.
Using this warm mixed refrigerant tailored to match the
cooling curve down to the 46 to 51C level and then a cold
mixed refrigerant to meet the refrigeration requirements below
that level allows this process to efficiently produce the desired
LNG stream.
Fig. 16-34 shows a simple schematic of one version of this
design, which has been used commercially in several locations
for base load liquefaction.
Precooled with Nitrogen Cycle
Liquefaction Process
APCI has developed the AP-X

liquefaction process, which


uses a third refrigerant cycle (nitrogen) to unload a propane
pre-cooled mixed refrigerant (C3MR) process. All of the 7.8 mil-
lion MTA LNG trains constructed in the last six years in Qatar
use this new hybrid process.
The front of the hybrid process as shown in Fig. 16-35 is es-
sentially identical to that of a conventional C3MR process. As in
that process, propane chills the inlet gas to around 40C, and
the mixed refrigerant cools the gas down to only 115C in a
spiral-wound heat exchanger (SWHE), instead of 151C typi-
cal of the C3MR process.
The nitrogen expansion cycle then serves as the subcooling
refrigerant, taking the liquefied gas temperature down the rest
of the way, cooling it to 151C. As with the C3MR process, the
rest of the temperature decrease occurs in the final pressure
reduction step.
The nitrogen cycle enables an increase in the trains through-
put at a given SWHE diameter. The SWHE diameter is what
normally limits the capacity of a liquefaction train using
SWHEs. By adding the nitrogen cycle, several refrigerant
streams are removed from the SWHE relative to the C3MR
process, making room for more tubes dedicated to feedstock
flow, and shortening the SWHE. The trade-off is the need for a
second SWHE in subcooling (nitrogen) service.
Turboexpander Based Process
Another route for commercialization not used in current on-
shore base load facilities is to use a liquefaction scheme based
on the use of isentropic expansion via a set of turboexpanders to
supply the refrigeration. A version of this is used in some peak
shaving applications where only a portion of the inlet gas has to
be liquefied.
This type of technology, while not as efficient as currently
used cascade and mixed refrigerant based processes has the ad-
vantage of avoiding or minimizing the refrigerant inventory re-
quired for the liquefaction step, and therefore, might be attrac-
tive for use offshore where large volumes of refrigerant would be
hazardous and lead to much greater mass and deck area
15
.
SOLIDS FORMATION
In addition to the obvious need for water removal from the
gas stream to protect from blockage in the cryogenic sections of
a plant, consideration must be given to the possible formation of
other solids or semi-solids in the gas stream. Amines, glycols,
and compressor lube oils in the gas stream can form blockages
in the system. Generally these contaminants will form a block-
age upstream of an expander, in the lower temperature ex-
change circuit, or on the screen ahead of an expander.
Carbon dioxide can form as a solid in lower temperature sys-
tems. Fig. 16-36 will provide a quick estimate for the possibility
of formation of solid CO
2
. If operating conditions are in the
methane liquid region as shown by the insert graph, the dashed
solid-liquid phase equilibrium line is used. For other conditions
the solid isobars define the approximate CO
2
vapor concentra-
tion limits.
For example, consider a pressure of 2400 kPa (abs). At 110C,
the insert graph (Fig. 16-36) shows the operating conditions to
be in the liquid phase region. The dashed solid-liquid phase
equilibrium line indicates that 2.1 mol percent CO
2
in the liquid
phase would be likely to form solids. However, at the same pres-
sure and 100C, conditions are in the vapor phase, and
1.28 mole percent CO
2
in the vapor could lead to solids forma-
tion. This chart represents an approximation of CO
2
solid for-
mation. Detailed calculations should be carried out if Fig. 16-33
indicates operation in a marginal range. This means that the
most probable condition for solid CO
2
formation may be several
trays below the top of the tower rather than at expander outlet
conditions. Again, while Fig. 16-36 indicates marginal safety
from solids formation, detailed calculations must be carried
out.
In addition to CO
2
and water which can solidify and cause
blockage and damage in cryogenic equipment, hydrocarbons
can also solidify at temperatures found in LNG plants. Figure
16-34 shows some freezing point temperatures for pure com-
pounds which can be troublesome in LNG facilities.
While all of the compounds listed can solidify at LNG tem-
peratures, the solubility of these compounds in LNG are such
that only at certain concentrations will there be solid formation.
Cyclohexane and benzene are the compounds with the highest
freezing points in this list. Cyclohexane is normally not present
in significant quantity in produced gas, however; benzene is
present in most gases and can be found in level in the 1000 ppm
range, well above the solubility limit at LNG temperatures.
Toluene can also be a problem, so typically aromatic compounds
are the highest concern due to the combination of concentration
present in the gas and freezing point.
16-29
FIG. 16-35
AP-X

LNG Process
FIG. 16-34
Typical Dual Mixed Refrigerant Process
16-30
FIG.16-36
Approximate Solid CO
2
Formation Conditions
16-31
FIG. 16-37
Solubility of Benzene in Methane
16-32
FIG. 16-38
Solubility of Benzene in Ethane
16-33
The solubility of these compounds in LNG streams is compo-
sition dependent. Fig. 16-37 shows the solubility of benzene in
methane, Fig.16-38 shows the solubility of benzene in ethane.
Comparison of these two figures shows that at say 162C, the
solubility in methane is about 2 ppm while the solubility in eth-
ane is 75 ppm. Heavier hydrocarbons such as propane and bu-
tane have even higher solubility numbers. Thus the composi-
tion of the LNG is an important factor in the solid formation
concentration of this compound and other components in Fig.
16-39. Typically, reduction of the benzene concentration to 10
ppm is sufficient to prevent solid formation. The GPA performed
research in this area (See Section 1) and has produced a predic-
tive computer program to calculate freezing points for both hy-
drocarbons and CO
2
in LNG streams.
NITROGEN REJECTION
Virtually all natural gas contains some amount of nitrogen
whVirtually all natural gas contains some amount of nitrogen
which lowers the heating value of the gas, but is so low to be no
particular problem. However, in some reservoirs gas contains
larger amounts of nitrogen than cannot be tolerated due to con-
tractual considerations on minimum heating content. In these
cases, the operator has three options:
1. Blend the gas with richer gas to maintain overall heating
value;
2. Accept a reduced price for sales gas or a less secure mar-
ket; or
3. Remove the nitrogen to meet sales specifications. Options
1 and 2 are reasonable approaches to the problem but are
very location specific.
When a nitrogen rejection unit (NRU) is selected as a pro-
cess option for a gas stream, it may be combined with NGL re-
covery in an integrated plant design. A block flow diagram of a
combined NGL/NRU facility is shown in Fig. 16-40.
16
The over-
all objective of this facility is to produce a nitrogen vent stream
with minimum hydrocarbon content (that is normally sent to
atmosphere), specification sales gas stream, and a specification
NGL product. A primary contributor to facility cost is the re-
quired gas compression. Regardless of the technology, recom-
pression of the sales gas is usually required unless the residue
gas can be marketed at 2050 kPa (ga) or less. Inlet compres-
sion is necessary if the gas is available at much less than 2750
kPa (ga) and can be justified for higher pressure gas depending
on the nitrogen rejection technology.
Cryogenic Technology
Nitrogen rejection is typically carried out using cryogenic
distillation technology to achieve very high hydrocarbon recov-
ery and minimize methane losses in the nitrogen vent stream.
Due to the low temperature operation the gas, after optionally
being compressed to required inlet pressure, is fed to a pretreat-
ment unit for CO
2
and water removal. The CO
2
will start freez-
ing at 57C, and therefore must be removed. For some technol-
ogy the CO
2
must be removed to 30 ppmv but modern designs
remove part of the hydrocarbon product at warmer tempera-
tures in a prefractionator and can tolerate higher CO
2
levels.
Typically, CO
2
removal is accomplished with amine treating
which can easily remove CO
2
to acceptable levels. See Section
21 for details concerning CO
2
removal. The dehydration step is
carried out with molecular sieve dehydration, which is covered
in Section 20. Another impurity requiring pretreatment to re-
move is mercury, which attacks the aluminum heat exchangers
in the low temperature section. Mercury removal techniques
are covered in Section 21. Typically, removal is accomplished
with an adsorbent bed.
The process design of the nitrogen rejection unit is a strong
function of the feed gas nitrogen content and should consider
the sales gas compression power requirement to give an opti-
mized design. For nitrogen contents below about 20%, a single
column with a heat pump cycle has been used such as shown in
Fig. 16-41. By operating the column at 2400 kPa (ga), low pres-
sure methane can be used in the condenser. The drawback to
this process is the heat pump compressor that is required to
minimize overall power consumption. The heat pump system
includes one or more intercondensers on the column resulting
in increased plant complexity. At nitrogen contents above about
30% an interlinked two column system such as in Fig. 16-42 is
the preferred choice. This design uses the nitrogen content of
the feed gas as column reflux and is efficient due to the con-
denser/reboiler arrangement. Lower feed gas nitrogen content
leads to an increased loss of methane in the column overheads
unless this basic process is modified. The high pressure (HP)
column operates at about 2400 kPa (ga). This design is quite
flexible and can be used at nitrogen contents above 50%. A re-
cycle compressor can be added to handle nitrogen contents be-
low 20% though this adds to power consumption. To reduce
power consumption and processing cost for feed gas with low
nitrogen content, prefractionator designs were developed
whereby fractionation is performed in a first, relatively high
pressure column to produce a low nitrogen content bottoms
stream, with a heat pump system or an interlinked two column
system downstream to effect almost total removal of hydrocar-
FIG. 16-39
Problem Compounds in LNG
Freezing Temp, C
Cyclohexane 6.55
Benzene 5.53
n-Decane 29.63
n-Nonane 53.48
n-Octane 56.76
n-Heptane 90.55
Toluene 94.98
n-Hexane 95.31
FIG. 16-40
Nitrogen Rejection Flow Diagram
14
Inlet
Gas
Inlet
Compression
Sales Gas
4825 kPa (ga)
Sales Gas
Compression
NGL
NGL / NRU
4825 kPa (ga) N2 Vent
Dehydration
CO2
Removal
CO2
16-34
bon from vented nitrogen. Newer licensed designs use a prefrac-
tionator with a downstream column at intermediate pressure
and use the evaporation of product hydrocarbon at one or more
pressure levels to minimize overall compression duty.
Early NRU designs typically produced all the hydrocarbon
product at 690 kPa (ga) or less whereas newer designs can pro-
duce up to about half the hydrocarbon at feed gas pressure and
the balance at intermediate pressures. These improvements
have significantly reduced power requirements, capital cost
and operating costs.
Variable nitrogen content feed gas requires careful design to
ensure efficient operation over a range of compositions. In en-
hanced oil recovery (EOR) applications, where the feed gas ni-
trogen content can range from less than 5% to about 80%, an
interlinked two column process may be optimal. However, the
heat pump based process can produce nitrogen at 2400 kPa
(ga), so a tradeoff study is usually required in EOR applications
to decide which process is the optimum one to use.
The NGL recovery section may be designed for ethane and
heavier recovery or propane and heavier recovery. Since NGL
recovery is also a cryogenic process, it is possible to integrate it
with the nitrogen rejection process to reduce capital and operat-
ing cost compared to separate NGL extraction and NRU facili-
ties. The NGL recovery is a traditional turboexpander setup
except that the front-end heat exchange is integrated with the
nitrogen and sales gas streams from the NRU section. The in-
cremental cost for NGL recovery may be quite small, because
many of the required process steps such as dehydration and
compression are already present although the operation can be
significantly more complex with multiple side streams feeding
the sales gas recompressor.
As such integrated plants need to be carefully designed for
variations in feed gas compositions to ensure they can meet
both NGL product and sales gas specifications with minimal
hydrocarbon content in the nitrogen stream.
Recovery Efficiencies
In the separation of nitrogen from natural gas, high purity
products are readily achievable. Sales gas purity of 2% or lower
nitrogen is achievable. The hydrocarbon content of the nitrogen
vent stream is typically specified at about 1% maximum to min-
imize environmental issues. This is equivalent to over 99.95%
hydrocarbon recovery for a low nitrogen content feed gas. NGL
recovery efficiencies associated with an integrated NGL/NRU
can be quite high as well. Ethane recovery of well over 90% with
virtually complete propane and heavier recovery is typical. If
ethane recovery is not desired, the process can be designed for
high C
3
+ recovery and incidental ethane recovery.
Molecular Sieves
Engelhard developed the Molecular Gate

technology using
titanium silicate materials whose pore size can be adjusted to
+/ 0.1 Angstrom. This technology is currently available for li-
cense from Guild Associates. These materials can be used to
preferentially adsorb molecules on a size exclusion basis. For
example, a material with 3.7 Angstrom pore diameter will ad-
sorb nitrogen with a molecular diameter of 3.6 Angstrom but
pass methane with a molecular diameter of around 3.8 Ang-
stroms. These units run best with an inlet pressure of 690
kPa (ga), but can operate from 480 to 900 kPa (ga).
FIG. 16-41
Single-Column NRU
16
FIG. 16-42
Two-Column NRU
16
16-35
The Molecular Gate

system utilizes a pressure swing


adsorption (PSA) process. It can also be designed for the si-
multaneous removal of CO
2
and nitrogen from natural gases
containing both impurities.
17
ENHANCED OIL RECOVERY
In order to increase oil production in many reservoirs, the
injection of gas for enhanced oil recovery (EOR) has been car-
ried out in numerous projects. The gas injection plan can lead to
three different types of processing facilities. First, high meth-
ane or high nitrogen gas can be injected for pressure mainte-
nance of the reservoir. In this case the gas is in a separate phase
from the oil phase, and any gas produced is simply recycled to
the reservoir. Processing of the gas in traditional gas processing
facilities is often carried out.
Second, the gas injected may be nitrogen with little or no
hydrocarbons. In this case the injection conditions are chosen
such that the nitrogen becomes miscible with the oil phase. As
the oil is produced, the nitrogen and associated gas are pro-
duced as a mixed gas phase. This produced gas can be reinjected
or processed for fuel, sales gas and NGL production. The pro-
cesses used separating the nitrogen in this case are described in
the Nitrogen Rejection subsection previously covered.
The third type of EOR process involves the injection of CO
2
.
Large volumes of CO
2
are injected into the reservoir and be-
come miscible with the oil phase. This CO
2
essentially scrubs
the oil from the reservoir and can greatly increase oil produc-
tion. As with the miscible nitrogen injection projects, the CO
2
is
produced with the oil and gas and must be handled in the gas
processing facilities. The CO
2
that is injected into the reservoir
is typically purchased from third party suppliers and is the
single greatest operating cost in the EOR project. Therefore, the
CO
2
produced with the associated gas is valuable and must be
recovered and recycled to the reservoir.
CO
2
Processing for EOR
The CO
2
produced in an EOR project can be separated from
the hydrocarbon components using solvent or membrane pro-
cesses as described in Section 21 of this Data Book. However,
solvent processes such as amines, potassium carbonate, and
physical solvents, as well as membrane systems, were not de-
signed to handle the large volumes of CO
2
, which are present in
the EOR gas. The capital and operating costs of these systems
increase in proportion to the acid gas content. Additionally, the
CO
2
is produced at low pressure and typically saturated with
water. The EOR project needs high pressure, dry CO
2
for rein-
jection.
An EOR gas processing plant is designed for three primary
separations. First, the methane in the gas is needed for fuel and
possibly to sell for additional revenue. Second, the produced gas
often contains hydrogen sulfide (H
2
S) which is removed from
the CO
2
stream for safety considerations. Third, EOR gas is
typically rich in recoverable NGLs. Fig. 16-43 is an example
EOR production profile. This example shows the effect of the
EOR operations on the gas to be handled. The CO
2
may start
out at a few percent but eventually builds to over 90% as the gas
volume increases. The NGL curve in Fig. 16-44 (on a CO
2
free
basis) shows that the hydrocarbon portion of the gas gets con-
tinually richer. In fact, in most projects, the in-situ oil is actu-
ally stripped of the midrange hydrocarbons such that over 10%
of the crude production is in the gas phase with the CO
2
.
FIG. 16-43
Example EOR Production Forecast
18
16-36
All of the required separations could be performed in a frac-
tionation process which would produce dry CO
2
at elevated
pressure as one of the products. Each step of the separation of
C
1
, CO
2
, H
2
S, and C
2
+ components has technology issues which
must be addressed with non-traditional concepts to achieve the
necessary separations by fractionation.
Separation of CO
2
and Methane
The relative volatility of CO
2
and methane at typical operat-
ing pressures is quite high, usually about 5 to 1. From this
standpoint, distillative separation should be quite easy. How-
ever, at processing conditions, the CO
2
will form a solid phase if
the distillation is carried out to the point of producing high pu-
rity methane. The phase equilibria considerations in this sepa-
ration are discussed in detail in Section 25 of this Data Book.
Fig. 16-44 illustrates the theoretical limits of methane pu-
rity, which can be obtained in a binary CO
2
/methane system. In
practice the purity limits of the methane product are around
15% CO
2
.
One approach to solving this methane-CO
2
distillation prob-
lem is to use an extractive distillation approach known as the
Ryan/Holmes process. This concept involves adding a heavier
hydrocarbon stream to the condenser in a fractionation column.
The addition of this stream, which can contain ethane and
heavier hydrocarbons, significantly alters the solubility charac-
teristics of the system such that virtually any purity of methane
can be produced.
Fig. 16-45 illustrates the effect of adding a third component
(in this case n-butane) to a CO
2
-methane distillation column
producing 2% CO
2
overhead. By adding n-butane, a column op-
eration profile without CO
2
solid formation can be achieved.
Adding greater amounts of the additive increases the safety
margin away from the CO
2
solid formation region. Other char-
acteristics of this additive addition concept include:
Raising the operating temperature of the overhead
Increasing CO
2
/methane relative volatility
Permitting higher pressure operation by raising the mix-
ture critical pressure
In practice the additive has been increased to the point that
propane refrigeration can be used for the overhead condenser
rather than cascade refrigeration. Under these conditions the
design acts much like a refrigerated lean oil process for NGL
recovery.
CO
2
-Ethane Separation
The separation of CO
2
and ethane by distillation is limited
by the azeotrope formation between these components. An azeo-
tropic composition of approximately 67% CO
2
, 33% ethane is
formed at virtually any pressure.
Fig. 16-46 shows the CO
2
-ethane system at two different
pressures. The binary is a minimum boiling azeotrope at both
pressures with a composition of about two thirds CO
2
and one
third ethane. Thus, any attempt to separate CO
2
and ethane to
nearly pure components by distillation cannot be achieved by
traditional methods. Extractive distillation is required.
As developed by Ryan and Holmes, the technique involves
the addition of a heavier hydrocarbon, usually butane or heavi-
er, to the top section of the distillation column.
FIG. 16-44
Distillation Profile CH
4
CO
2
Binary
19
FIG. 16-45
Distillation Profile Binary Feed with nC
4
Additive
20
16-37
The upper dashed line in Fig. 16-46 represents the phase
behavior of a multicomponent feed distilled with a butane-plus
additive. With this technique, virtually any purity of CO
2
and
ethane is thermodynamically possible.
For the CO
2
-methane separation, the additive is introduced
in the condenser. In the CO
2
-ethane separation, the additive is
normally introduced several trays below the top of the
column. The primary CO
2
-ethane distillation is achieved below
the additive feed tray. It is in this area that the relative volatil-
ity of the CO
2
to ethane is reversed to remain above 1.0 and the
azeotrope is circumvented. High relative volatilities are ob-
tained at all points on and below the additive feed tray.
In the top portion of the column above the additive feed tray,
no resolution of the azeotrope is achieved, as the relative vola-
tility of CO
2
/ethane is less than 1.0. This part of the column
serves as a recovery zone for the extractive distillation addi-
tive.
While the separation of these species is feasible as described,
the value of the ethane may not justify extracting it from the
recycled CO
2
.
Separation of CO
2
and H
2
S
The distillative separation of CO
2
and H
2
S can be performed
with traditional methods. The relative volatility of CO
2
and H
2
S
is quite small. While an azeotrope between H
2
S and CO
2
does
not exist, the vapor liquid equilibrium behavior for this binary
approaches azeotropic character at high CO
2
concentrations.
21
In many cases the CO
2
is required to contain less than 100
ppmv H
2
S. In order to achieve such purity a very large fraction-
ation tower is required with large energy requirements.
Another aspect to be considered is the CO
2
in the bottom
(H
2
S concentrated) stream. If fed to a Claus sulfur recovery
plant, the CO
2
/H
2
S ratio is desired to be less than 2 to 1. Achiev-
ing such a low ratio will require high energy input in many
cases.
By adding a third component, as in the CO
2
-ethane separa-
tion system, the relative volatility of CO
2
to H
2
S is significantly
enhanced. Fig. 16-47 demonstrates the relative volatility en-
hancement due to addition of n-butane to the CO
2
H
2
S binary
system.
Thus, if a system containing CO
2
, ethane and H
2
S is pro-
cessed in an extractive distillation column, the ethane and H
2
S
can be separated from the CO
2
. The exact specification for pu-
rity and recovery will determine the system operating require-
ments. From a thermodynamic standpoint the CO
2
could be
produced overhead with the ethane and the H
2
S (and any C
3
+
components) produced as a mixed bottom product.
FIG. 16-46
Vapor-Liquid Equilibria CO
2
C
2
H
6
20
FIG. 16-47
CO
2
H
2
SnC
4
System at 4100 kPa
22
16-38
Again this separation may not be economically attractive,
especially since there are H
2
S selective solvents available for
extracting H
2
S, while leaving the bulk of the CO
2
. If this route
is elected, then this process should be done on the raw gas prior
to dehydration, since these solvents all contain water as a con-
stituent. See Section 21 for more details.
Overall Process Configuration
The EOR processing steps can be arranged in a system to
achieve all the desired separations although the process con-
figuration can take on several variations, the configuration
most often used in EOR processing plant is shown in Fig. 16-
48.
In this configuration, the first step is the ethane/CO
2
sepa-
ration in the ethane recovery column. This separation is carried
out at pressures in the 2400 kPa (ga) range using refrigeration
for reflux in the 18C range. The CO
2
and lighter components
are taken overhead, compressed to around 4500 kPa (ga) and
sent to the CO
2
recovery column. This is a bulk removal column
which produces CO
2
as a liquid bottom product. This CO
2
can
then be pumped to reinjection. The overhead product is essen-
tially a CO
2
/C
1
binary which is limited by CO
2
solid formation
considerations. This binary stream is then separated by use of
the extractive distillation step to produce a methane stream
with low CO
2
content.
The bottoms products from the ethane recovery and demeth-
anizer columns are combined and processed in the additive re-
covery column. In this column the additive, which is a C
4
+
stream, is separated from the lighter hydrocarbons for recycle
to the distillation columns. A net C
4
+ product is also produced.
The additive used in the ethane recovery and demethanizer col-
umns is continuously regenerated and reused much the same
as lean oil in traditional gas processing applications. The dis-
tinct difference in this case is that this C
4
+ stream is generated
from the feed gas and is used as an extractive distillation agent
rather than as an absorption agent.
The light NGL product produced overhead in the additive
recovery column also contains any H
2
S which was present in
the feed gas and some residual CO
2
. This product is usually
treated in a small amine or dry bed unit to meet sales specifica-
tions. The acid gas may then be sent to a sulfur recovery unit.
This four column EOR processing plant is designed to han-
dle the wide range of feed rates and compositions encountered
in EOR projects. In the design effort early, peak and late year
cases must be investigated to ensure proper operation over
time. This is especially important since the exact timing, flow
rate and composition of EOR production are extremely difficult
to predict. CO
2
breakthrough to the processing plant can occur
rapidly. In some projects the CO
2
volume can triple in less than
one year. As the EOR process has matured, other configura-
tions have been developed which mix technologies such as
membranes with Ryan/Holmes facilities to optimize the capital
and operating costs over the project life.
FIG. 16-48
Four-Column Ryan/Holmes Process
18
16-39
REFERENCES
1. Rusten, B.H., Gjertsen, L.H., Solbraa E., Kirkerd T., Toril Hau-
gum T. and Puntervold S., Determination Of The Phase Enve-
lope Crucial For Process Design And Problem Solving, 87th
GPA Annual Convention, Grapevine, TX , March 2-5, 2008.
2. Maddox, R.N., and Erbar, J.H., Low-Pressure Retrograde Condensa-
tion, Oil & Gas Journal, July 1977.
3. Petroleum Extension Service, Field Handling of Natural Gas,
Third Edition, University of Texas, Austin, TX, 1972, pp. 36-51.
4. Sachin, J. and Lokhandwala, K., Heavy or Sour Fuel Gas? Max-
imize Uptime Using a Simple Membrane Process to Improve Gas
Quality, 2010 Gas Machinery Conference, Phoenix, AZ, October
4-6, 2010.
5. Toreja, J., VanNostrand, B., Chan, N. and Dickinson, J.P., Ro-
tary-valve, Fast-cycle Pressure-swing Adsorption Technology Al-
lows West Coast Platform to Meet Tight California Specifications
and Recover Stranded Gas, 61
st
Laurence Reid Gas Condition-
ing Conference, Norman, OK, February 2011.
6. Russell, T.H., Truck-Mounted Plant Speeds Gas Processing, Oil &
Gas Journal, February 1, 1982, pp. 129-133.
7. Russell, T. H., Straight Refrigeration Still Offers Processing
Flexibility, Oil & Gas Journal, January 24, 1977.
8. Petroleum Extension Service, Plant Processing of Natural Gas,
University of Texas, Austin, TX, 1974, pp. 19-36.
9. Pitman, R.N., Hudson, H. M. and Wilkinson, J.D, Next Genera-
tion Processes for NGL/LPG Recovery, 77th GPA Annual Con-
vention, Dallas, TX, March 16, 1998.
10. Wilkinson, J.D. and Hudson, H.M., Improved NGL Recovery,
71st GPA Annual Convention, Anaheim, CA, March 1992.
11. Nasir, P., Sweet, W., Elliot, D., Chen, R. and Lee, R.J., Enhanced
NGL Recovery Process
sm
Selected for Neptune Gas Plant Expan-
sion, 82n
d
GPA Annual Convention, March 9-12, 2003.
12. Ulmanns Encyclopedia of Industrial Chemistry, VCH Publish-
ers, New York, NY, 1991, Vol 17, pp. 100-109.
13. Price, B.C. and Mortko, R.A., PRICO A Simple, Flexible Proven
Approach to Natural Gas Liquefaction, Gastech 96, Vienna, Aus-
tria, December 3-6, 1996.
14. Chatterjee, N, Kinard, G.E., and Geist, J.M., Maximizing Pro-
duction in Propane Precooled Mixed Refrigerant LNG Plants,
Seventh Conference on Liquefied Natural Gas, Jakarta, Indonesia,
May 15-19, 1983.
15. Finn, A. J., Floating LNG Plants Scaleup of Familiar Technol-
ogy, 88th GPA Annual Convention, San Antonio, TX, March
2009.
16. Price, B.C., NRUs Upgrade Production, Cut Costs, American
Oil and Gas Reporter, March 1994, pp. 56-60.
17. Mitariten, M.,Dolan, W. , and Maglio, A., Innovative Molecular Gate
Systems for Nitrogen Rejection Carbon Dioxide Removal and
NGL Recovery, 81st GPA Annual Convention, Dallas, Texas,
March 2002.
18. Price, B.C., Looking at CO2 Recovery in Enhanced Oil Recovery
Projects, Oil & Gas Journal, December 24, 1984, pp. 48-53.
19. Nagahana, K., Kobishi, H., Hoshino, D., and Hirata, M., Binary Va-
por-Liquid Equilibria of Carbon Dioxide-Light Hydrocarbons at
Low Temperatures, J. Chem. Eng. Japan 7, No. 5, p. 323 (1974).
20. Holmes, A.S., Ryan, J.M., Price, B.C., and Styring, R.E., Pilot
Tests Prove Out Cryogenic Acid-Gas/Hydrocarbon Separation
Processes, 61st Annual GPA Convention, Dallas, TX, March 15-17,
1982.
21. Sobocinski, D.P., Kurata, F., Heterogeneous Phase Equilibria of
The Hydrogen Sulfide-Carbon Dioxide System, AIChEJ.,5, No.4, p.
545 (1959).
22. Ryan, J.M., and Holmes, A. S., Distillation Separation of Carbon Di-
oxide from Hydrogen Sulfide, U.S. Patent No. 4,383,841 (1983).
16-40
NOTES:

You might also like