You are on page 1of 9

Ind. Eng. Chem. Res.

2006, 45, 337-345 337

GENERAL RESEARCH

Development of a Detailed Gasoline Composition-Based Octane Model


Prasenjeet Ghosh,* Karlton J. Hickey,† and Stephen B. Jaffe‡
Compositional Modeling Group, ExxonMobil Process Research Laboratories, Paulsboro Technical Center,
Paulsboro, New Jersey 08066

We present a model that predicts the research and motor octane numbers of a wide variety of gasoline process
streams and their blends including oxygenates based on detailed composition. The octane number is correlated
to a total of 57 hydrocarbon lumps measured by gas chromatography. The model is applicable to any gasoline
fuel regardless of the refining process it originates from. It is based on the analysis of 1471 gasoline fuels
from different naphtha process streams such as reformates, cat-naphthas, alkylates, isomerates, straight runs,
and various hydroprocessed naphthas. Blends of these individual process streams are also considered in this
work. The model predicts the octane number within a standard error of 1 number for both the research and
motor octane numbers.

1. Introduction Lovell et al.4 were an early group to identify that aromatics


and branched i-paraffins have higher ONs than the correspond-
Octane number (ON) is one of the most important properties ing paraffins. In the middle 1950s, the American Petroleum
of gasoline streams and is a measure of its antiknock property. Institute (API) Research Project 455-7 analyzed the pure-
It is defined as the volume percentage of i-octane in a blend of component ONs for over 300 different hydrocarbon molecules,
n-heptane and i-octane, which produces the same knock intensity and several reliable correlations relating gasoline composition
as the test fuel under standard test conditions in an ASTM to ON were developed. The work not only quantified the ON
internal combustion engine. ASTM defines two different types trends with molecular structure and size, it also studied the
of ONs, the research octane number (RON) and the motor octane
nonlinear interactions between different molecular types toward
number (MON), which are evaluated using the ASTM D2699
ON.7 However, the work was primarily focused on pure-
and the ASTM D2700 tests, respectively.1,2 Both methods use
component studies with a limited number of binary or ternary
the same standard test engine but differ in the operating
gasoline blends. Commercial gasoline contains many different
conditions. RON is measured in an engine running at 600 rpm
molecules in the C4-C13 range, and the more recent papers in
and a fuel/air mixture at a temperature of 60 °F, while MON is
measured with the engine running at 900 rpm and a fuel/air the literature have therefore focused on predicting the octane
mixture at a temperature of 300 °F. The slower engine speed number of such multicomponent mixtures. Anderson et al.8
and the lower fuel/air temperature as required in the RON test developed a useful method for predicting the RON of different
are representative of the fuel performance for city driving, while gasolines based on the gas chromatographic (GC) analysis of
the faster engine speeds and higher fuel/air temperature represent the sample. A total of 31 molecular lumps were considered to
the fuel performance for highway driving. describe the composition of gasoline, and the contribution of
Knock results from the premature combustion of the gasoline each lump was added linearly to compute the octane number
due to compression in the engine.3 As the fuel/air mixture is of the fuel. Although simple in its structure, the method by
compressed in the internal combustion engine, certain molecules Anderson et al.8 is rather restrictive in its use and is known to
in gasoline tend to self-ignite even before they reach the ignition show an error of around 2.8 numbers on average for catalytically
spark, thereby creating a resistive expansive motion in the cracked naphthas.9 Part of the less than satisfactory predictions
compression stroke of the engine and hence the knock. Depend- is perhaps the assumption of linearity in octane blending because
ing on the thermal stability of the molecule (which depends on octane blending is known to exhibit nonlinear interactions (both
its molecular structure) and the ensuing radicals, certain synergistic and antagonistic) among the various constituent
molecules tend to combust sooner (and knock more) than others. hydrocarbon molecular classes (e.g., paraffins, olefins, aromatics,
Consequently, ON is a direct function of the molecular etc.).7 Sasano10 described a procedure similar to that of Anderson
composition of the gasoline fuel, and any modeling effort should et al.8 to predict ON of the gasoline fuel from its composition
explicitly acknowledge it. measured by GC. The ON of the fuel is calculated as the linear
Numerous studies in the past have attempted to mathemati- volumetric average of the ON of each molecule in the fuel,
cally describe the ON as a function of the gasoline composition. which, in turn, is calculated using a linear correlative model.
Van Leeuwen et al.11 used nonlinear regression methods,
specifically projection pursuit regression and neural networks,
* To whom correspondence should be addressed. E-mail:
prasenjeet.ghosh@exxonmobil.com. to correlate the gasoline composition from GC to ON. However,
† E-mail: karlton.j.hickey@exxonmobil.com. both these techniques require much fine-tuning and experience
‡ E-mail: stephen.b.jaffe@exxonmobil.com. to develop. This is especially true for neural network based
10.1021/ie050811h CCC: $33.50 © 2006 American Chemical Society
Published on Web 11/24/2005
338 Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

approaches, where the determination of the number of layers


in the network, number of nodes in each layer, the kind of
transfer functions for each node, etc., is highly user-dependent.
Also, the lack of an underlying phenomenological structure and
the inherent opacity in such models make them less reliable in
the extrapolative mode. More recently, Meusinger and Moros
used genetic algorithms and neural networks to quantify the
partial ON of each gasoline component in the mixture based
on the structural elements of the molecule.12,13 Other relevant
work in this area includes the work by Lugo et al.,14 Twu and
Coon,15 and Albahri,16 who have all employed different varia-
tions of the correlating gasoline composition to its octane.
Although the literature is replete with many relevant papers,
most of the work falls into one of the two following catego-
ries: first, predicting the ON of individual hydrocarbon
molecules in gasoline by correlating molecular structure descrip- Figure 1. Schematic showing the difference between octane numbers of
tors such as topological indices, length of the backbone carbon the gasoline sample (ON), pure-component octane numbers (ONA and ONB),
chain, degree of branching, type of carbon atom based on and the corresponding blend octane numbers (BON ON
A and BB , respectively)
spectroscopy, etc., to the pure-component ON and, second, for a typical binary blend. Notice that the blend numbers are obtained by
predicting the ON of the actual gasoline fuel by correlating the extrapolating the tangent to the octane curve to the pure-component limit.
molecular composition of the gasoline fuel to its ON. In this
different hydrocarbon molecules, each contributing to the ON
work, we are interested in accomplishing the second objective.
of the gasoline fuel. Let ON denote the measured octane number
In principle, any model that achieves the second objective should
for the gasoline fuel while ONi represents the pure-component
also be able to predict (or reconcile) the ON of the individual
octane number for each molecule i in the fuel. Because a
hydrocarbon molecule because a pure component is merely the
molecule i may not necessarily always contribute its pure-
limiting value of a mixture. However, most of these published
component octane number to the gasoline fuel, each molecule’s
mixture models do not achieve this. Part of this lack of
contribution toward the fuel octane number is quantified by its
reconciliation in the published models originates from their
purely empirical and statistical nature of the underlying math- blend value, denoted by BON i . The blend value of a molecule
ematical structure used to develop the models, making them depends on the overall composition of the fuel it is part of.
rather restrictive for extrapolation. Further, most of the published Figure 1 schematizes the difference between ONi and BON i . By
models have focused on developing correlations that work well definition, ON is a linear volumetric blend of the blend
with individual or a very few naphtha process streams col- contributions of all of the different molecules present in the
lectively like fluidized catalytically cracked (FCC) naphtha, gasoline fuel. Therefore
reformates, or some light straight-run (LSR) naphtha streams.
The streams were so chosen that the ONs varied in a narrow ON ) ∑i ViBON
i (1)
range between 80 and 95 numbers. Although this has tradition-
ally been the most important range for commercial gasoline where Vi is the volume fraction of molecule i in the sample.
fuels, it would be beneficial to have a model that can work Experimental studies in our laboratory over many years have
with a much extended range of ONs between 30 and 120 revealed that the blend value of a molecule i varies almost
numbers. This is especially critical in today’s refinery blending linearly with the ON of the gasoline fuel it is part of, or blended
operations, where many different blend components (i.e., into. These results have been reported previously in refs 17 and
naphthas from different processes) with widely different ONs 18. A typical result highlighting this observation is shown in
need to be economically blended. Finally, none of the above Figure 2.18 The figure plots the variation of the blend values of
work predict the octane number of gasoline fuels that contain various n-paraffins (nC5-nC9) against the ON of the three
oxygenates such as methyl tert-butyl ether (MTBE), ethanol different refinery naphtha fuels in which they were volume-
(EtOH), or tert-amyl methyl ether (TAME), which are finding blended into. For each of the paraffins, the figure shows that
increasing use in commercial gasoline. the blending value of n-paraffin increases linearly as the ON of
We attempt to address some of the questions raised in the the refinery naphtha fuel increases. We have observed similar
previous paragraph. Specifically, our objectives are 2-fold: first, results over many other such studies. Thus, in general, we may
to develop a predictive model for ON, both RON and MON postulate that the blend value of molecule i may be ap-
for any gasoline fuel, dependent solely on the composition and proximated as a linear function of the gasoline ON. Therefore
independent of the refining process stream and, second, to extend
the range of applicability of the model to a much broader range ) a(0)
BON i + ai (ON)
(1)
i (2)
of octanes from 30 to 120 numbers. In developing such a model,
we would like to ensure that, in the limiting case of a pure Equation 2 is analogous to the equation for any partial molar
component, the model predicts the pure-component octane
property, where the partial molar property (BON i ) depends on
number. Further, we would also like to extend the model to
the composition of the gasoline fuel it blends into. BON i is
fuels containing different oxygenates. The details of such a
model development are presented in the next section. parametrized by two parameters, a slope a(1) i and an intercept
a(0)
i . Although the variation of the blend value of the individual

2. Development of the Octane Model molecules with the ON of the gasoline fuel is taken to be linear,
this does not imply a linear relationship of the fuel composition
Each gasoline fuel, regardless of the process stream (e.g., to its ON. Using the definition for ON from eq 1 in eq 2, it is
reformate, FCC, LSR, etc.), is a complex mixture of many easy to see that eq 2 captures multicomponent interaction.
Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 339

Figure 3. Typical nonlinear interactions between two species x1 and x2.


For instance, x1 could be a paraffin while x2 could be a olefin. Curve a
Figure 2. Variation of blending ON of n-paraffins in different refinery indicates positive interaction, curve b indicates no interaction, and curve c
naphtha fuels. indicates negative interaction.

Although eq 2 requires two parameters, a(0) i and a(1)


i , for of the gasoline fuel. A contribution is considered beneficial if
each molecule, it is possible to eliminate one of these parameters the blend value of the molecule/lump (BON i ) is greater than its
using the special case when the gasoline fuel is a pure pure-component octane number (ONi). Likewise, the contribu-
component. For instance, if the gasoline fuel is pure toluene, tion is detrimental if BON < ONi. For the two cases (a) βi < 1
i
then its ON is the same as the pure-component ON for toluene, and ONi < ON or (b) βi > 1 and ONi > ON, the lump i
which is the same as its blend value. Application of this contributes beneficially to the fuel ON. For the other two cases
boundary condition yields (a) βi > 1 and ONi < ON or (b) βi < 1 and ONi > ON, it
contributes detrimentally to the fuel ON. βi ) 1 is the special
ONi ) a(0)
i + ai (ONi) w ai ) (1 - ai )ONi
(1) (0) (1)
(3) case where the lump contributes equally to its pure-component
octane number.
If we define βi ) 1 - a(1)
i , eq 2 may be rewritten as Equation 5 forms the core of the octane model and is suitable
for predicting octanes of reformates, LSR, and alkylate streams.
i ) βiONi
a(0) (4) Although nonlinear in composition, the extent of nonlinearity
expressed in it is insufficient to fully capture the nonlinear
Rearranging eqs 1, 2, and 4, we get interaction between paraffins and olefins and/or paraffins and
naphthenes, which may be important for cat-naphtha fuels.
∑i ViβiONi lumps
∑ ViβiONi Published studies in the literature6,7 and independent research
in our laboratories reveal that hydrocarbons belonging to the
ON ) ) (5) same molecular class blend linearly; i.e., paraffins blend linearly
∑i Viβi ∑ Viβi
lumps
with other paraffins, olefins blend linearly with other olefins,
and so on. However, a blend of paraffins and olefins may exhibit
significant deviation from linearity. Such nonlinear interaction
This is the basic octane prediction model, where the summation in a binary blend is qualitatively described in Figure 3. Figure
index i runs over all of the molecules present in the gasoline 3a shows a positive interaction or equivalently a positive
fuel. However, a typical gasoline fuel contains an enormous deviation from linearity, Figure 3c shows a negative interaction,
number of different molecules, so some molecular lumping is and Figure 3b shows no interaction. Paraffins and naphthenes
necessary to build a practical model. Such lumping is reflected also exhibit similar deviations from linearity. In contrast, olefins
in the second part of eq 5, where we have replaced the summing and naphthenes tend to blend linearly. The blending behavior
variable i with “lumps” to indicate that the sum runs over all of aromatics as a group is somewhat unclear because of the
of the molecular lumps relevant for the octane model. A lump differences in the blending behavior of individual aromatic
defines the compositional level of abstraction that the model molecules and also because of the difficulty in measuring the
uses. It could be a single molecule like n-heptane or a group of high ONs of such blends. Also, note from Figure 3 that, for
similar molecules such as C10 aromatics. Note that, in light of certain blends, the ON of the blend may actually be higher or
this lumping, the term pure-component ON for lump i may not lower than both the individual ON extremes defined by the pure-
always necessarily reflect the ON of a single molecule. For component numbers. The mathematical structure of eq 5 is
molecular lumps that correspond to single molecules, ONi is incapable of describing this behavior because it will always be
the pure-component ON of that single molecule, but for bounded between the extremes of pure-component numbers for
molecular lumps that correspond to a group of similar molecules, the two components of the binary blend, as shown in eq 6.
ONi is the average of the pure-component ONs of all of the

( V1β1
) V2β2
( )
different molecules in that lump. These lumps are described in
more detail in section 3. ON1 e ON ) ON1 + ON2 e
βi’s are the adjustable parameters and represent whether a
V1β1 + V2β2 V1β1 + V2β2
molecule contributes beneficially or detrimentally to the ON ON2 (6)
340 Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

The nonlinear interaction of Figure 3 may be quantitatively


described by the following mathematical function for a binary
blend:

12 V2V1
k(a)
y) (7)
1 + k(b)
12 V2

where k(a) (b)


12 and k12 are interaction parameters. By appropriately
choosing these interaction parameters, we can describe positive,
negative, and no blending interactions. The structure has been
borrowed from the more familiar rate expression for inhibiting
and synergistic pyrolysis kinetics which tend to exhibit similar
behavior. For a ternary blend where there are multiple binary
interactions, i.e., between V1 and V2, between V1 and V3, and
likewise between V2 and V3, eq 7 may be written as
Figure 4. Pure-component RONs for various i-octane isomers.

12 V2V1
k(a) + 13 V3V1
k(a) + 23 V2V3
k(a)
y) (8) Using eqs 2-4 in eq 9 and rearranging the various terms, we
1+ 12 V2 +
k(b) 13 V3 +
k(b) 23 V3
k(b) get

Equation 8 has six interaction parameters: k(a) (b)


12 and k12 for the
∑ ViβiONi + IP∑P ViβiONi
PONA
1-2 interaction, k13 and k13 for the 1-3 interaction, and k(a)
(a) (b)
23 ON ) (11)
and k(b)
23 for the 2-3 interaction, respectively. It is easy to see
from eqs 7 and 8 that if nonlinear interaction has to be modeled ∑ Viβi + IP(∑P Viβi - ∑P Vi)
PONA
for every binary interaction in a typical gasoline fuel, it would
lead to a large number of interaction parameters; e.g., even for This is the complete octane model. Note that when I ) 0, eq
a very lumped description of a typical gasoline fuel with 50 11 reduces identically to eq 5. Also, in the limiting case when
different molecular lumps, this leads to 50C2 ) 1225 binary the gasoline fuel is a pure component (Vi ) 1), eq 11 yields
interactions where nCr ) n!/r!(n - r)!. Consequently, the ON ) ONi, thereby returning the pure-component octane
nonlinear interaction will be modeled only at the P/O/N/A level number.
(i.e., at the level of the total paraffins, total olefins, total
naphthenes, etc., and not at the individual molecule level). 3. Defining the Molecular Lumps
For the purposes of this paper, we will consider the nonlinear
Next we need to define the various molecular lumps that will
interaction between paraffins and naphthenes and between be used in the octane model. The molecular lumps have been
paraffins and olefins. Thus, a correction similar to eq 8 can now selected based on the following two criteria:
be added to eq 1 to model the nonlinear interactions in octane (a) Analytical Differentiation. We have defined separate
blending. Thus lumps where it is possible to analytically measure and dif-

( )
ferentiate the lump for most of the process streams. For instance,
PNVN + kPOVO
k(a) (a) it is possible to analytically measure the different aromatics by
ON ) ∑ ViBON
i + ∑P VjBON
j (9) carbon number for most of the naphtha process streams by
PONA 1+ PNVN +
k(b) POVO
k(b) different GC techniques; however, it is not always possible to
differentiate between the various aromatic isomers at each
carbon number with current analytical techniques. Consequently,
where VN and VO represent the total naphthenes and olefins in for the aromatics, our lump definition has been restricted to
the gasoline sample. Mathematically, VN ) ∑i∈naphthenesVi and total aromatics by carbon number only.
VO ) ∑i∈olefinsVi. Note that in eq 9, while the first sum runs over (b) ON Differentiation. When analytical differentiation is
P, O, N, and A (i.e., paraffins, olefins, naphthenes, and possible across similar molecular lumps, we have defined
aromatics) molecules, the second sum runs only over the separate lumps only if their ONs are widely different and their
paraffinic molecules. If the interaction term is denoted by I, eq relative distributions differ across various process streams. For
9 may be more succinctly rewritten as example, we consider separate lumps for mono-, di-, and
trimethyl-i-paraffins at each carbon number as opposed to a
ON ) ∑
PONA
(1 + Ii)ViBON
i (10) single lump for the total i-paraffins at each carbon number. Such
delumping is necessary for two reasons: first, the ONs of the
branched i-paraffins vary widely with the degree of branching
and, second, the relative distribution of branched isomers is very
where

[( ]
different across different naphtha streams. Figure 4 plots the

)
pure-component ONs of all of the i-octane isomers and shows
PNVN + kPOVO
k(a) (a)
that, depending upon the particular isomer of i-octane, whether
for i ∈ P it is a mono-, di-, or trimethyl isomer, its ON could vary between
Ii ) 1 + k(b)
PNVN + kPOVO
(b)
25 and 100 numbers. The relative distribution of these isomers
0 otherwise is also different in different naphtha process streams, as shown
Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 341

Table 1. Typical Measured Isomer Distribution (wt %) for MONs for most of the lumps have been chosen from the values
i-Octanes reported in the API-45 project.5,6 For lumps that correspond to
stream type monomethyls dimethyls trimethyls individual molecules, e.g., n-butane, n-pentane, etc., the pure-
alkylates 0 18.5 81.5 component values are taken as is from ref 6, while for lumps
reformates 68.9 31.1 0 that correspond to a group of similar molecules such as
monomethyl-i-octanes, an average value is chosen based on the
in Table 1 for alkylates and reformates. Trimethyl-i-octanes
ONs of all of the different monomethyl-i-octane isomers. Two
dominate C8 i-paraffins in alkylates, whereas monomethyl-i-
important trends emerge from this dataset: first, ON decreases
octanes dominate C8 i-paraffins in reformates. The observations
with the carbon number and, second, ON increases with the
reported for i-octane are also true for other i-paraffins. Conse-
degree of branching at the same carbon number. Such trends
quently, i-paraffins are modeled as three lumps at each carbon
are shown for the paraffins and aromatics in Figures 5 and 6,
number in the model based on their degree of branching. The
respectively, where the pure-component ONs are plotted against
situation is somewhat different for aromatics or olefins. At each
the number of carbon atoms in the hydrocarbon molecule. These
carbon number, the variation in the RONs for the aromatic
trends have been used to estimate the pure-component ONs of
isomers is only between 5 and 10 ONs (see Figure 6), which is
the lumps that are not reported in ref 6. The final set of pure-
not significant enough to introduce noticeable errors in the
component RONs and MONs for the various molecular lumps
predictions across the various process streams, especially
in the model is shown in Table 2.
because the relative abundance of the individual isomers is small
and their distribution does not change appreciably either across
4. Experimental Program
the process streams. The situation is similar for i-olefins, where
the ON does not change significantly depending on whether An extensive database of 1471 gasoline fuels was collected
the particular i-olefin is a mono-, di-, or trimethylolefin. In from many naphtha process streams found in the petroleum
addition, there is great difficulty in analytical speciation of these refinery. These include 143 alkylates, 165 cat-naphtha (FCC)
individual isomers for most of the process streams. For streams gasolines, 440 reformates, 366 hydroprocessed naphtha streams,
where such detailed information might be available, e.g., 117 SCANfining [Selective CAtalytic Naphtha Refining, a
aromatic isomer distribution in reformates, the fact that the proprietary ExxonMobil Technology] products, 40 LSR naph-
isomers are always at equilibrium makes the delumping un- thas, 13 isomerates, and 187 finished blends of these different
necessary as a single lump with an average ON of the lump process streams. Fuels with oxygenates contained TAME,
(reconciled with the equilibrium isomer distribution) would be MTBE, and EtOH in the range of 2-10 vol % oxygenates. Each
adequate. of these 1471 fuels was analyzed for its detailed composition
On the basis of the above two criteria, we have considered a using a multitude of different GC techniques. RON and MON
total of 57 compositional lumps in this work to describe any were also measured on these fuels. Because the composition
naphtha stream. These include 32 lumps for n- and i-paraffins, was measured through multiple GC columns, each measuring
6 lumps for naphthenes, 7 lumps for aromatics, 9 lumps for certain carbon number ranges of the composition, a thorough
olefins and cyclic olefins, and 3 lumps for oxygenates. The data-reconciliation procedure was employed to reconcile the
lumps range from individual molecules to a group of similar analysis by the different analytical measurements. For example,
molecules at various levels of abstraction. Although the oxygen- fuel samples with both the overall PIONA analysis and the
ates are ON improvers, they are treated in the model similar to specific GC analysis (which yield detailed information in a
all of the hydrocarbon lumps. These lumps are summarized in particular carbon number range) were checked to ensure that
Table 2. the two analyses were consistent. Likewise, the compositional
Once the lumps have been defined, the next step is to specify information based on GC and PIONA was reconciled with the
their pure-component ONs. The pure-component RONs and boiling point curve wherever available. Further, many repeats
Table 2. Various Lumps Considered in the Present Octane Model along with Their Pure-Component RONs and MONs
RON MON RON MON RON MON
Paraffins Paraffins (cont’d.) Aromatics (cont’d.)
n-butane 94 89.6 n-undecane -35 -35 C12 aromatics 102 90
isobutane 102 97.6 C11 monomethyl 5 5
n-pentane 62 62.6 C11 dimethyls 35 35 Olefins/Cyclic Olefins
i-pentane 92 90.3 C11 trimethyls 90 82 n-butenes 98.7 82.1
n-hexane 24.8 26 n-dodecane -40 -40 n-pentenes 90 77.2
C6 monomethyls 76 73.9 C12 monomethyl 5 5 i-pentenes 103 82
2,2-dimethylbutane 91.8 93.4 C12 dimethyls 30 30 cyclopentene 93.3 69.7
2,3-dimethylbutane 105.8 94.3 C12 trimethyls 85 80 n-hexenes 90 80
n-heptane 0 0 i-hexenes 100 83
C7 monomethyls 52 52 Naphthenes total C6 cyclic olefins 95 80
C7 dimethyls 93.76 90 cyclopentane 100 84.9 total C7d 90 78
2,2,3-trimethylbutane 112.8 101.32 cyclohexane 82.5 77.2 total C8d 90 77
n-octane -15 -20 m-cyclopentane 91.3 80
C8 monomethyls 25 32.3 C7 naphthenes 82.0 77 Oxygenates
C8 dimethyls 69 74.5 C8 naphthenes 55 50 MTBE 115.2 97.2
C8 trimethyls 105 98.8 C9 naphthenes 35 30 TAME 115 98
n-nonane -20 -20 EtOH 108 92.9
C9 monomethyls 15 22.3 Aromatics
C9 dimethyls 50 60 benzene 102.7 105
C9 trimethyls 100 93 toluene 118 103.5
n-decane -30 -30 C8 aromatics 112 105
C10 monomethyls 10 10 C9 aromatics 110 101
C10 dimethyls 40 40 C10 aromatics 109 98
C10 trimethyls 95 87 C11 aromatics 105 94
342 Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

Figure 5. Variation of pure-component RONs for various paraffins (adapted from ref 5).

Figure 6. Variation of pure-component RONs for various aromatics (adapted from ref 5).

were run on the RON and MON measurements of each fuel to an upper bound of 10. Although, in principle, β may be <0,
ensure high fidelity in the dataset. the restriction of β g 0 was necessary to avoid singularity in
eq 11. Softer constraints in the form of inequality constraints
5. Results were also employed to ensure that β did not vary arbitrarily
with the carbon number within the same molecular class but
5.1. Parameter Estimation: Constrained Least-Squares rather maintained the directional trend with the carbon number
Formulation. A constrained least-squares minimization problem and branching as shown previously in Figures 5 and 6.
was solved using the Levenberg-Marquadt algorithm in order The experimental dataset was partitioned into a training set
to regress the parameters of the model. The adjustable param- (90% of the samples) and a testing set (10% of the samples),
eters β were allowed to vary between a lower bound of 0 and and many such partitions (total partitions considered ) 200)
Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 343

Figure 7. Model predictions for RONs for various process streams and blends. N ) number of samples. SE ) standard error.

Table 3. Parameter Values


molecular class molecular lumps β(RON) β(MON)
n-paraffins nC4-nC12 2.0559 0.3092
i-paraffins C4-C12 mono-, di-, and trimethyl-i-paraffins 2.0204 0.4278
naphthenes C5-C9 naphthenes 1.6870 0.2821
aromatics benzene-C12 aromatics 3.3984 0.4773
olefins/cyclic olefins C4-C12 linear, branched, and cyclic olefins 8.9390 10.0000
oxygenates MTBE, EtOH, TAME 3.9743 2.0727
interaction parameters k(a) (b) (a) (b)
PN, kPN, kPO, kPO
0.2, 2.4, 0.4, 3.6 0.2, 2.4, 0.4, 3.6

Table 4. Model Performance across Various Process Streams


were considered to ensure the robustness of the parameter
estimates. Many different initial guesses were assumed and the RON MON
parameters reoptimized to ensure that the optimization problem SE st dev SE st dev
was not trapped in an inferior local solution. Sensitivity analysis total 1.0194 1.0296 1.0572 1.0263
in the form of a local perturbation of the parameters was also alkylates 0.8822 0.9122 0.6935 0.6505
performed to see whether the parameters returned to the same cat-naphthas 0.6732 0.6065 0.5468 0.5846
optimal or not. reformates 0.8869 0.9115 0.9132 0.9889
hydroprocessed naphtha 1.2359 1.2352 1.1718 1.1094
Although the final set of parameter values (β) for all of the products
individual molecular lumps used in this model cannot be SCANfining products 1.2188 1.2278 1.1926 1.1033
disclosed in the publication for proprietary reasons, the average LSR naphtha 1.4531 1.3800 1.3921 1.2598
values of these parameters, grouped by P, I, O, N, and A classes, isomerates 0.7683 0.6734 0.7522 0.2977
are reported in Table 3. Also included in the table (last row) process blends 0.9095 0.9104 1.3255 1.2048
are the four interaction parameters used in the model. The results and blends. The individual SE and the standard deviation are
based on these average parameter values are satisfactory, though summarized in Table 4. The FCC gasoline fuels are predicted
we may point out that the use of individual β values for the with the highest accuracy (SE ) 0.67), while the predictions
different molecular lumps enhances model predictions. The on the straight-run fuels have the highest error (SE ) 1.45)
results presented below are based on the individual β parameters. relative to other process streams. The higher relative error is
5.2. Prediction Statistics. The results of the model are shown due to the higher measurement error in the ONs of straight-run
in Figures 7 and 8 for both RON and MON across the 1471 naphtha fuels, which have octanes in the range of 40-70
gasoline fuels. The plots summarize not only the overall numbers where the octane test itself is less than satisfactory.
performance of the model but also the performance across the The process blends, many of which have 2-10 vol % oxygen-
individual process streams. The overall standard errors (SEs) ates, are also predicted quantitatively, suggesting that the model
defined as xΣ(meas - pred)2/n for RON and MON are 1.01 structure is sufficient to capture the effect of oxygenates on the
and 1.05 numbers, respectively, on all of these process streams gasoline fuel. The prediction results on all of the process streams
344 Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006

Figure 8. Model predictions for MONs for various process streams and blends. N ) number of samples. SE ) standard error.

Table 5. Comparison of Model Predictions with and without Data


Reconciliation of the Measured Composition
SE (Standard Error)
model predictions model predictions
(using the measured (using autotuned
composition “as is”) compositions)
refinery 1 0.9105 0.8444
refinery 2 0.5457 0.4482
refinery 3 0.7016 0.7069
refinery 4 1.0197 1.0085
refinery 5 1.1686 1.1682
refinery 6 0.7850 0.7833
refinery 7 1.2074 0.8436
Figure 9. Error distribution for RON across the various samples.
many commercial gasoline fuels from seven different refineries
and their blends are within the measurement error; therefore, it worldwide was collected. The gasoline fuels collected were
is unlikely that the average error can be further reduced for a
blends of various refinery process streams such as alkylates,
model that spans such a diverse set of naphtha fuels.
reformates, FCC-naphtha, and so on. The individual process
An important requirement of any regression model is that its
streams are therefore the “blend components” of the final
predictions are unbiased with respect to its inputs. To ensure
gasoline blend. On each individual blend component, the
this, the frequency distribution of the model errors is plotted in
composition was measured using a combination of GCs along
Figure 9. The abscissa in the figure represents the error level in
with RON and MON. In addition, some bulk properties such
the model predictions, while the ordinate represents the number
of fuels predicted at these various error levels. Approximately as gravity and the boiling point curve were also measured. On
75% of the samples are predicted within (1.0 ON, 93% ()75 the final gasoline blends, only RON and MON were measured.
+ 18) are predicted within (2.0 ON, and 99% of the samples On the basis of the measured composition of the blend
are predicted within (3.0 ON. The frequency distribution of components and using the blend recipe (i.e., the volumetric ratios
Figure 9 closely resembles a normal distribution, suggesting of how the components were blended), the composition of the
that the model is independent of any inherent bias. Further, the final gasoline blend was calculated and RON and MON were
plot of the prediction error against each of the compositional predicted using eq 11. A comparison of the model predictions
lumps also revealed no directional trend of the composition with against the measured RON for these fuels is shown in the first
the model error. column of Table 5. The results are segmented based on the
5.3. Independent Model Validation and Prediction Im- performance in each individual refinery. The table shows that,
provement via Data Reconciliation. To test the pure predictive on average, the model predicts all of the data with a SE of (0.9
capabilities of the model, an independent dataset comprising number, which is within the experimental error of the different
Ind. Eng. Chem. Res., Vol. 45, No. 1, 2006 345

measurements, thus providing credence to the robustness and in the model predictions are demonstrated using a data-
reliability of the model. Similar results are obtained for MON reconciliation algorithm used in tandem with the predictive
as well. model.
However, a part of the error in the model predictions
originates from the error in measuring the composition itself Acknowledgment
and the elimination of which might improve the model predic- We express our appreciation to N. L. Avery, D. I. Hoel, M.
tions. We will use a data-reconciliation formulation, which we R. Apelian, K. D. Rose, R. J. Quann, and C. R. Kennedy for
call autotuning, to eliminate this error contribution. Autotuning, their many helpful discussions and suggestions.
described in eq 12, attempts to make minimal changes in the
measured composition subject to the constraints that target Literature Cited
measured properties are matched exactly or very closely
approximated within an  tolerance, where  is a very small (1) Knock characteristics of motor fuels by the research method.
American Society for Testing and Materials Test Method D2699; American
number, g0. Society for Testing and Materials: West Conshohocken, PA, 1989.

( )
(2) Knock characteristics of motor fuels by the motor method. American
w - wmeasured 2 Society for Testing and Materials Test Method D2700; American Society
min for Testing and Materials: West Conshohocken, PA, 1989.
w wmeasured (3) Cornelius, W.; Caplan, J. D. Some Effects of Fuel Structure,
s.t. Tetraethyl Lead and Engine Deposits on Precombustion Reactions in a Firing
Engine. SAE Quart. Trans. 1952, 6 (3), 488.
Pj -  e Pj(w) e Pj +  (4) Lovell, W. G.; Campbell, J. M.; Boyd, T. A. Detonation Charac-
∀ j ) 1, ..., m (12) teristics of Some Aliphatic Olefin Hydrocarbons. Ind. Eng. Chem. 1931,
∑ wi ) 1 23, 555.
(5) Knocking characteristics of pure hydrocarbon. Special Technical
wg0 Publication No. 225; American Society for Testing and Materials: West
Conshohocken, PA, 1958.
In eq 12, wmeasured is the vector of raw measured composition (6) API Technical Data Book on Petroleum Refining; API: Washington,
DC, 1986.
from the GC, w the autotuned composition reconciled to a (7) Scott, E. J. Y. Knock Characteristics of Hydrocarbon Mixtures. Proc.
number of different measured properties, and Pj the various API DiV. Refin. 1958, 38, 90.
target measured properties. Because the formulation of eq 12 (8) Anderson, P. C.; Sharkey, J. M.; Walsh, R. P. Calculation of Research
results in a highly underdetermined set of equations, the choice Octane Number of Motor Gasolines from Chromatographic Data and a New
Approach to Motor Gasoline Quality Control. J. Inst. Pet. 1972, 59, 83.
of the number and type of property constraints is important to (9) Huskey, D.; Zhrmann, U. Determinacion de Octanaje RON mediante
obtain a better autotuned composition. The following property Cromatografia de Gases en Columnas Capilares. Informe Tecnico InteVep;
constraints were used in this work: (a) RON, (b) MON, (c) INTEVEP: Caracas, Venezuela, 1988.
gravity at 60 °F, (d) total paraffin/i-paraffin ratio, (e) total (10) Sasano, Y. Measuring method of research octane number of gasoline
olefins, and (f) boiling point curve. The results based on the by gas chromatograph and its apparatus. JP Patent 09-318613, 1997.
(11) Van Leeuwen, J. A.; Jonker, R. J.; Gill, R. Octane number prediction
autotuned compositions of the blend components are shown in based on gas chromatographic analysis with non-linear regression tech-
the second column of Table 5. Remarkably, using an autotuned niques. Chem. Intell. Lab. Syst. 1994, 24, 325.
composition consistently improves the ON predictions for the (12) Meusinger, R.; Moros, R. Determination of quantitative structure-
gasoline blends across all of the seven refineries. On average, octane rating relationships of hydrocarbons by genetic algorithms. Chem.
Intell. Lab. Syst. 1999, 46, 67.
the SE improved from (0.9 numbers to (0.8 numbers. This (13) Meusinger, R.; Moros, R. Determination of octane numbers of
improvement may seem marginal, but even a 0.1 number gasoline compounds from their chemical structure by 13C NMR spectroscopy
improvement in the prediction error below a (1 number has a and neural networks. Fuel 2001, 80, 613.
significant economic consequence. (14) Lugo, H. J.; Ragone, G.; Zambrano, J. Correlations between Octane
Numbers and Catalytic Cracking Naphtha Composition. Ind. Eng. Chem.
Res. 1999, 38, 2171.
6. Conclusions (15) Twu, C. H.; Coon, J. E. Estimate octane numbers using an enhanced
method. Hydrocarbon Process. 1997, 76, 65.
We have developed a composition-based predictive model (16) Albahri, T. A. Structural Group Contribution Method for Predicting
for both RON and MON that can be universally applied across the Octane Number of Pure Hydrocarbon Liquids. Ind. Eng. Chem. Res.
a wide variety of gasoline fuels derived from different naphtha 2003, 42, 657.
process streams and blends. Each gasoline fuel is composition- (17) Jaffe, S. B. Control of Gasoline Manufacture. U.S. Patent 4,251,-
870, 1981.
ally represented by 57 different molecular lumps and by a (18) Heck, R. H. Contribution of Normal Paraffins to the Octane Pool.
combination of different GCs and correlated to the ON. The Energy Fuels 1989, 3, 109.
model structure permits a wide range of composition extrapola-
tion from pure components to real gasoline blends. Its predic- ReceiVed for reView July 11, 2005
tions are within a SE of (1 number for both RON and MON ReVised manuscript receiVed October 10, 2005
Accepted October 14, 2005
across the multitude of gasoline fuels. The model is applicable
for a broad range of ONs from 30 to 120. Further improvements IE050811H

You might also like