5-95 gap chart

244
REFINERY-WIDE OPTIMIZATION by XUAN LI, B.E. A DISSERTATION IN CHEMICAL ENGINEERING Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY Approved May, 2000

Upload: processpipingdesign

Post on 01-Nov-2014

101 views

Category:

Documents


7 download

DESCRIPTION

astm 5-95

TRANSCRIPT

Page 1: 5-95 gap chart

REFINERY-WIDE OPTIMIZATION

by

XUAN LI, B.E.

A DISSERTATION

IN

CHEMICAL ENGINEERING

Submitted to the Graduate Faculty of Texas Tech University in

Partial Fulfillment of the Requirements for

the Degree of

DOCTOR OF PHILOSOPHY

Approved

May, 2000

Page 2: 5-95 gap chart

COPYRIGHT 2000, XUAN LI

Page 3: 5-95 gap chart

ACKNOWLEDGEMENTS

Refinery-wide optimization would become "Mission Impossible" without

innumerable help from other people. Many people have given help in many different

ways. It is imperative for me to give them my appreciation.

I want to express my sincere appreciation to Dr. James B. Riggs for providing me

with the opportunity to study at Texas Tech University, for guiding me and encouraging

me throughout the whole project, and for supporting me financially during the course of

this work. His style of clarifying every statement and developing insight of every

problem and his determination to get things done will influence me in my whole career.

I also want to extend my sincere thanks to Dr. Theodore F. Wiesner for his

guidance in my qualifying project. His fighting spirit and positive attitude show me the

correct way to deal with obstacles. I would also like to thank my other committee

members. Dr. Richard W. Tock and Dr. W. J. Bryan Oldham for their assistance in my

Ph.D. research. Appreciation is extended to the industrial members of the Texas Tech

Process Control Consortium for their financial support and invaluable inputs to my

research. A special thanks go to Scott Boyden of Aspen Technology Incorporation and

Dr. Charlie Cutler, for their input on my research.

This work would not have much industrial relevance without the plant data

obtained from the refinery considered in this work. I wish to express my appreciation to

all the engineers and technicians who helped me during their busy schedule.

I also wish to extend my thanks to Dr. Taskar and Robert Ellis for leading me

through the modeling and optimization of reformer and FCCU and for introducing the

NFL to me. Go Cowboys! I would also like to thank Joe, Marshall, and Matt, for their

technical support on all computer issues. Special thanks go to Govindhakannan for

reviewing my dissertation draft and for innumerous discussion on research and life,

which is invaluable during the last one and a half years.

Page 4: 5-95 gap chart

Finally, I would like to thank my parents for their love and support through the

years. Lastly, I would like to dedicate this work to my dear wife, Jenny. Let us be

together and start a more exciting life!

ni

Page 5: 5-95 gap chart

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ii

ABSTRACT vii

LIST OF TABLES ix

LIST OF FIGURES xiv

CHAPTER

1. INTRODUCTION 1

2. LITERATURE REVIEW 8

2.1 Single-Unit Modeling and Feed Characterization 8

2.2 Refinery-Wide Optimization 14

3. CRUDE UNIT MODELING 16

3.1 Introduction 16

3.2 Atmospheric Tower Modeling 20

3.2.1 Feed Characterization 22

3.2.2 Side-Draw Product Calculation 28

3.2.3 Atmospheric Tower Furnace Calculation 36

3.2.4 Flash Zone Calculation 36

3.2.5 Side Stripper Calculation 42

3.2.6 Draw Tray Locations 43

3.2.7 Draw Tray Temperature Calculation 45

3.2.8 Side-Draw product temperatvire calculation 48

3.3 Vacuum Tower Modeling 50

3.3.1 Vacuum Tower Furnace Calculation 51

3.3.2 Side-Draw Product Calculation 52

3.3.3 Flash Zone Calculation 52

3.4 Separation Section Calculation 54

3.5 Comparison between the Developed Model and Tray-to-Tray ChemCad Model 55

3.6 Auxiliary Processes in the Crude Unit 63

3.6.1 Rerun Unit 64

IV

Page 6: 5-95 gap chart

3.6.2 Residue Oil Solvent Extraction (ROSE) Unit 65

3.6.3 Debutanizer 66

3.6.4 Naphtha Splitter 66

4. FCC FEED CHARACTERIZATION AND MODEL BENCHMARKING 68

4.1 Process Overview 68

4.2 Model IV FCC Unit Modeling 71

4.3 FCC Feed Characterization 73

4.3.1 Volume and Weight of FCC Feed 73

4.3.2 FCC feed Characterization 74

4.3.3 Weight Fractions ofLumps in the FCC Feed 82

4.4 Model Benchmarking 87

4.5 FCC Gasoline Octane Model Modification 92

5. REFORMER FEED CHARACTERIZATION AND MODEL BENCHMARKING 94

5.1 Process Overview 94

5.2 Reformer Modeling 96

5.3 Reformer Feed Characterization 100

5.3.1 Naphtha Desulfurizer 100

5.3.2 Paraffins-Naphthenes-Aromatics (PNA) in the Reformer Feed 103

5.3.3 Molar Flow Rates of the Pseudo-Components in the Reformer Feed 108

5.4 Reformer Model Benchmarking 110

5.5 Average Reformer Operation 114

5.5.1 Operation Time Fraction of Low Severity 114

5.5.2 On-Stream Factor 115

5.5.3 Regeneration Cost 118

6. MODELING OF GAS PLANT, ALKYLATION UNIT AND DIESEL HYDROTREATER 120

6.1 Gas Plant 120

6.1.1 Sources of Light Gas 121

6.1.2 Fuel Gas Production 123

6.1.3 Depropanizer 124

Page 7: 5-95 gap chart

6.2 Alkylation Unit 126

6.3 Diesel Hydrotreater 134

7. GASOLINE BLENDING MODELING 136

7.1 Process Overview 136

7.2 Properties of the Gasoline Blending Stocks 141

7.2.1 Light Straight-Run (LSR) Gasoline 141

7.2.2 Other Gasoline Blending Stocks 145

7.3 Gasoline Blending Model 148

7.3.1 Octane Model 148

7.3.2 RVP Blending Model 152

7.3.3 Percent Distilled Model 155

8. REFINERY-WIDE OPTIMIZATION 156

8.1 Formulation of the Optimization Problem 156

8.1.1 Formulation of the Obj ecti ve Function 15 6

8.1.2 Obj ective Function Evaluation 15 9

8.1.3 Price Structure 160

8.1.4 Decision Variables 162

8.1.5 Constraints 166

8.1.6 Optimization Algorithm 171

8.2 Optimization Case Studies 171

8.2.1 Base Case 171

8.2.2 Summer Mode 172

8.2.3 Winter Mode 181

8.2.4 Optimal Solution Analysis 187

8.2.5 Profitability Improvement 189

8.3 Single-Unit Optimization 190

9. CONCLUSIONS AND RECOMMENDATIONS 196

9.1 Conclusions 196

9.2 Recommendations 200

BIBLIOGRAPHY 203

APPENDIX 208

VI

Page 8: 5-95 gap chart

ABSTRACT

A fuel-oriented refinery converts crude oil into various fuel products which are

used for transportation and heating. It also provides feedstock for petrochemical plant.

Based on a real refinery in the Gulf coast, a first-principle, nonlinear, plant-wide

model was developed by integrating several single-unit models into an overall model.

Detailed models were developed for two major units, crude unit and gasoline blending.

The crude unit model is a non-stage-by-stage, steady-state model based on material

balance and energy balance. The model calculates the yields and properties of the

products based on the feed information and product specifications. The gasoline blending

model calculates the complete set of gasoline specifications of three grades of gasoline

from the information of gasoline blending stocks.

Existing detailed models of fluidized catalytic cracking (FCC) unit and catalytic

reformer were used in this work after minor modification. Simplified first-principle

models were developed for other units in the refinery. Each single-unit model was

benchmarked against the industrial data obtained from the refinery.

In the refinery-wide model, the outputs of models of upstream units are used as

the inputs to the models of downstream units. The intermediate streams are characterized

in the overall model to provide necessary information for the models of downstream

units. Detailed composition information of feeds to FCC unit and catalytic reformer is

calculated. General properties, volume, weight, and specific gravity are calculated for

other intermediate streams.

A constrained nonlinear optimization was carried out using the developed

refinery-wide model. The objective of the optimization is to maximize the daily revenue

of the whole refinery. The decision variables are the collection of the process variables of

each unit that has significant influence on the economy of whole refinery operation. The

nonlinear and linear constraints in the optimization are the summation of constraints of

each unit. Two operation modes. Summer Mode and Winter Mode, were studied. The

optimal solutions obtained from refinery-wide optimation show that the revenue increase

vu

Page 9: 5-95 gap chart

over the normal operating conditions is 4.5% for Summer Mode and 3.6%) for Winter

Mode. The revenue from refinery-wide optimization is about 1.6% over single-unit

optimization.

vni

Page 10: 5-95 gap chart

LIST OF TABLES

3.1 Feedstock Makeups. 22

3.2 Gap (5-95) ASTM Temperature between Adjacent Side-Draw Products in the Atmospheric Tower. 29

3.3 A and B Variables for Equation 3.10. 33

3.4 Pressure in the Atmospheric Tower. 37

3.5 Tray Numbers of Separation Sections in Atmospheric Tower. 44

3.6 Draw Tray Location. 44

3.7 Separation Specifications of Side-Draw Products. 57

3.8 Comparison of the Gains Obtained from the Simplified Model with ChemCad model, ASTM 95% Point, Furnace Outlet Temperatiire, and VGO TBP End Point versus Product Flow Rate, Summer Mode. 60

3.9 Comparison between the Gains Obtained from the Simplified Model and ChemCad Model, Pumparound Heat Duty versus Product Flow Rate, Atmospheric Tower, Summer Mode. 61

3.10 Comparison between the Gains Obtained from the Simplified Model and ChemCad Model, Pumparound Return Temperature versus Product Flow

Rate, Vacuum Tower, Summer Mode. 62

4.1 Boiling Range ofLumps in Ten-Lump FCC Reaction Network. 73

4.2 Industrial Data of FCC Feed. 75 4.3 Comparison between the Industrial Data and Feed Characterization of the FCC

Feed. 82

4.4 Weight Fractions of Eight Lumps in an FCC Feed of API=23. 84

4.5 Weight Fractions of the Eight Lumps in the FCC feed across the Operating Range. 86

4.6 FCC Model Benchmarking: Industrial Data and Model Prediction. 88

4.7 Adjustable Variables in FCC Model Benchmarking. 89

4.8 Process Constraints of FCC Unit. 91

4.9 FCC Octane Model. 93

5.1 Octane Ratings of the Typical Gasoline Blending Stocks. 94

5.2 Chemical Components ofthe Reformer Feed. 98

5.3 Ratios in the Calculation of the Naphtha Hydrotreater. 101

IX

Page 11: 5-95 gap chart

5.4 Industrial Data of Crude A. 104

5.5 Industrial Data of Crude B. 104

5.6 Industrial Data of Crude C. 105

5.7 Industrial Data of Crude D. 105

5.8 Comparison of Volume Percentages of PNA from Feed Characterization and

the Industrial Data. 108

5.9 Catalyst Weight in Each Reactor Bed. 110

5.10 Comparison ofthe Industrial Data and Model Prediction after Benchmarking. 111

5.11 Adjustable Prameters in the Reformer Model Benchmarking. 112

5.12 Reformer Operation Limit. 116

5.13 Coke Contents in Dfferent Regions. 116 5.14 Ratios between Average Coking Rate and Coking Rate at the Begiiming of a

Cycle. 117

6.1 Light Gas Production Rates in the Crude Unit. 122

6.2 Compositions of Light Gas from FCC Unit. 123

6.3 Losses of Heavy Hydrocarbons in the Fuel Gas. 124

6.4 Portion ofthe Propylene in the C3 Product. 124

6.5 Weight Percentage ofthe Propylene that Reacts in the Second Reaction Path

in the Alkylation Unit. 129

6.6 Octane Number ofthe Butylene Alkylate and Propylene Alkylate. 130

6.7 Weight Percentage ofthe Unreacted Isobutane in the Alkylation Unit. 132

6.8 Weight Ratios ofthe Hydrogen and the Products in the Diesel Hydrotreater. 134 6.9 Volumetric Ratios ofthe Hydrogen and the Products in the Diesel

Hydrotreater. 135

7.1 Volumes and Sources ofthe Gasoline Blending Stocks in Typical Summer Mode Operation. 137

7.2 Specification of Three Grades of Gasoline Produced in the Refinery

Considered in This Work. 138

7.3 Industrial Data ofCrude A and Crude B. 142

7.4 Industrial Data of Crude C and Crude D. 143

7.5 Properties ofGasoline Blending Stocks. 147

7.6 Universal Set ofthe Binary Interaction Parameter between Components. 150

8.1 Chemicals Purchased by the Refinery Considered in This Work. 157

Page 12: 5-95 gap chart

8.2 Price Structure ofthe Refinery-Wide Optimization. 158

8.3 Decision Variables of Refinery-Wide Optimization. 164

8.4 Nonlinear Constraints of Refinery-wide Optimization. 167

8.5 Comparison of Model Prediction ofthe Base Case with the LP report. Summer Mode. 173

8.6 Comparison of Model Prediction ofthe Base Case with the LP Report, Winter Mode. 174

8.7 Optimum Values ofthe Decision Variables of Refinery-wide Optimization,

Summer Mode. 177

8.8 Active Constraints in Refinery-wide Optimization, Summer Mode. 178

8.9 Comparison ofthe Decision Variables of Refinery-wide Optimization with Base Case, Summer Mode. 179

8.10 Comparison ofthe Decision Variables of Refinery-wide Optimization with

Base Case, Winter Mode. 183

8.11 Active Constraints in Refinery-wide Optimization, Winter Mode. 184

8.12 Comparison ofthe Product Slates ofthe Optimum Solution with the Base

Case, Winter Mode. 185

8.13 The Mean and Variance of the Feasible Solutions. 187

8.14 Change of the Objective Function Value around Optimal Solutions. 189

8.15 Profitability Improvement of the Refinery-wide Optimization. 190

8.16 Prices of the Side-draw Products from the Crude Unit, Summer Mode. 191

8.17 Prices of the Side-draw Products from the Crude Unit, Summer Mode. 192 8.18 Comparison of Single-unit Optimization with Refinery-wide Optimization,

Summer Mode. 193 A.l Constants of Polynomial Expression for the TBP Curve ofthe Mixed Crude in

Summer Mode, Calculating the Vol.%) Given the Temperature in °F. 208

A.2 Constants of Polynomial Expression for the TBP Curve ofthe Mixed Crude in Summer Mode, Calculating the Temperature in °F Given the Vol.%. 208

A.3 Constants of Polynomial Expression for the TBP Curve ofthe Crude in Winter Mode, Calculating the Vol.% Given the Temperature in °F. 209

A.4 Constants of Polynomial Expression for the TBP Curve ofthe Crude in Winter Mode, Calculating the Temperature in °F Given the Vol.%. 209

A.5 Constants of Polynomial Expression for the API Gravity Curve ofthe Mixed Crude in Summer Mode, Calculating the API Gravity Given the Vol.%. 210

XI

Page 13: 5-95 gap chart

A.6 Constants of Polynomial Expression for the API Gravity Curve ofthe Mixed Crude in Winter Mode, Calculating the API Gravity Given the Vol.%. 210

A.7 Constants of Polynomial Expression for the Sulfur Curve ofthe Mixed Crude in Summer Mode, Calculating the Sulfur Given the Vol.%o. 211

A. 8 Constants of Polynomial Expression for the Sulfur Curve ofthe Mixed Crude in Winter Mode, Calculating the Sulftir Given the Vol.%. 211

A.9 Constants of Polynomial Expression for Converting ASTM End Point to TBP End Point. 212

A. 10 Constants of Polynomial Expression for Converting Gap (5-95) ASTM to Gap (0-100) TBP. 212

A. 11 Constants of Polynomial Expression for Calculating the Molecular Weight of a Crude Cut Given the Mean Average Boiling Point and API gravity. Characteristic Factor from 12.1 to 12.6. 213

A. 12 Constants of Polynomial Expression for calculating the molecular weight of a Crude Cut Given the Mean Average Boiling Point and API gravity. Characteristic Factor from 11.7 to 12.0. 213

A. 13 Constants of Polynomial Expression for Calculating the Molecular Weight of a Crude Cut Given the Mean Average Boiling Point and API gravity. Characteristic Factor from 11.3 to 11.6. 214

A. 14 Constants of Polynomial Expression for Calculating the Enthalpy of a Crude Cut from Its API Gravity, Temperature and Phase. 214

A. 15 Constants of Polynomial Expression for Calculating the Enthalpy of a Crude Cut in Saturated Vapor Phase from Its API Gravity and Temperature. 219

A. 16 Constants of Polynomial Expression for Calculating the Steam-free Delta T Given the Value of the Percent Stripout of Crude Cuts. 221

A. 17 Constants of Polynomial Expression for Calculating Stripout from Steam Rate. 222

A. 18 Constants of Polynomial Expression for Calculating the Steam-free DT Minus Actual DT(F) from the Temperature (F) Difference between Feed and Stripping Steam and Percent Stripout. 223

A. 19 Constants of Polynomial Expression for Calculating the Enthalpy of Steam from Temperature. 223

A.20 Constants of Polynomial Expression for Calculating the Enthalpy of Air from Temperature. 224

A.21 Constants of Polynomial Expression for Calculating the Enthalpy of Water from Temperature. 224

xu

Page 14: 5-95 gap chart

A.22 Constants of Polynomial Expression for Calculating the Slope of Flash Reference Line from Corresponding Distillation Reference Line. 225

A.23 Constants of Polynomial Expression for Calculating the Ratio of DT(flash)/DT(TBP) from Volumetric Percent Distillated. 225

A.24 Constants of Polynomial Expression for Calculating the Temperature Difference between the Distillation and Flash Reference Curves from the Slope of Distillation Reference Curve. 226

A.25 Constants of Polynomial Expression for Calculating the T50 of Flash Curve under Vacuum from the Pressure and T50 ofthe Flash Curve under Atmospheric Pressure. 227

xni

Page 15: 5-95 gap chart

LIST OF FIGURES

3.1 Block Flow Diagram of a Fuel-Oriented Refinery. 17

3.2 Schematic of a Fuel Type Crude Unit. 18

3.3 Schematic of the Model of the Atmospheric Tower of the Crude Unit. 21

3.4 Crude Cuts on a TBP-Volumetric Curve. 23

3.5 TBP Curve of the Crude Used in Summer Mode. 25

3.6 API Curve of the Crude Feed Used in Summer Mode. 27

3.7 Sulfur Curve ofthe Crude Feed Used in Summer Mode. 27

3.8 Flash Zone of the Atmospheric Tower. 39

3.9 Flowchart for the Calculation of Flash Zone Pressure and Temperature. 40

3.10 Energy and Material Balance Quantities at a Side-Draw Tray. 46

3.11 Flowchart ofthe Side-Draw Tray Temperature Calculation. 49

3.12 Schematic of a ChemCad Crude Unit Model. 56

3.13 Comparison of Simplified Model and ChemCad Rigorous Model, Summer Mode. 58

3.14 Comparison of Simplified Model and ChemCad Rigorous Model, Winter

Mode. 58

3.15 Comparison of Temperature Profile in the Atmospheric Tower. 63

4.1 Schematic of a Model IV Fluidized Catalytic Cracking (FCC) Unit. 70

4.2 Ten-Lump Reaction Network. 72

5.1 Process Flow Scheme of a Semi-Regenerative Catalytic Naphtha Reformer. 97

6.1 Schematic of the Gas Plant in the Fuel-Oriented Refinery. 121

6.2 Schematicof a Hydrofluoric Acid Alkylation Unit. 127

7.1 Flowchart for the Calculation ofthe Tio%, TSQO/O, T9OO/„ ofGasoline Blends. 154

8.1 Flowchart of the Execution of the Refinery-Wide Model. 161

8.2 Comparison of Product Slates of Refinery-Wide Optimization with Base Case, Summer Mode. 180

8.3 Comparison of Product Slates of Refinery-Wide Optimization with Base Case, Winter Mode. 186

8.4 Comparison of Product Slates of Single-Unit Optimization (Crude Unit) with Refinery-Wide Optimization, Summer Mode. 195

xiv

Page 16: 5-95 gap chart

CHAPTER 1

INTRODUCTION

Crude oil is the dominant fuel resource in the world today. Products from crude

oil are extensively used in industry and normal life. The products directly coming from

crude oil include fuel gas, liquidified petroleum gas (LPG), gasoline, jet fuel, diesel,

heating oils, lubricated oil, fuel oils, solvents, asphalt, etc. Crude oil is also the raw

material for the petrochemical and chemical industries.

To use crude oil efficiently and to make its usage environmental safe, it is

necessary to refine crude oil into various products that have different specifications that

satisfy the usage requirement and environmental regulations. The process of refining the

crude oil into various products is normally carried out in refineries. Although there is a

long history of using crude oil as burning material, it is not until 1860 that the first real

petroleum refinery was built at Titusivlle, PA, at a cost of about $15,000 (Nelson, 1958).

Since then, petroleum refining has developed into a major industry in almost every

country. In 1998, the total worldwide crude oil production was about 73 million barrels

per day. The worldwide refining capacity is about 67 million barrels per day (Beck,

1999). Population growth and continued world economic expansion produce an ever-

increasing demand for fuel. This indicates that worldwide refining capacity will keep

increasing in the future.

Although conventional fuel from petroleum faces challenges from fuels obtained

from renewable resources, it is expected to remain a dominant player at the beginning of

the 21st century (Bensabat, 1999). The worldwide oil reserves in 1999 are about I trillion

barrels. If the oil production is maintained at the level of 1998, the current oil reserve can

sustain about 39 years. With technical advancement and further exploration, the oil

reserve keeps increasing. The reasonable forecast is that crude oil will stay as the largest

fuel resource in the first half of 21st century.

Petroleum refining has developed from simple separation in the early stage to a

very complex process today. The early development of petroleum refining technology

Page 17: 5-95 gap chart

includes applying continuous distillation, vacuum distillation, and thermal cracking, etc.

The refining industry has undergone tremendous expansion and change since World War

II. Many new processes with high efficiency have been invented. These new processes

include Fluidized Catalytic Cracking (FCC), Catalytic reforming, Alkylation, catalytic

desulfurization, delayed coking, etc. Enormous increases in the size of process units, new

catalytic processes, shifting product demands, and new sources of petroleum from tar

sands and oil shales have made present-day technology and economics of petroleum

refining a very complex and sophisticated science (Gary and Handwerk, 1984).

For decades, the large and mature U.S. transportation fiiel market has been

dominating the global petroleum supply (Bensabat, 1999). Consequently, transportation

fuels including gasoline, jet fuel and diesel, has the largest quantity among all the

products made from crude oil. By far, passenger vehicles make up the largest sector with

respect to number of vehicle and fuel consumption. Most of U. S. passenger vehicles are

fueled by gasoline in conventional combustion engines. They consume about 8.5 million

barrel per day of gasoline, or 12% of global petroleum demand (Bensabat, 1999). Most of

transportation fuel consumed in U. S. is produced domestically. Due to such massive

production, many refineries in U. S. are fuel-oriented refineries whose main fiinction is to

produce transportation fiiel. There are other types of refineries. For example, some

refineries provide aromatics and olefin for petrochemical pleints. Some refineries also

produce lubricants and asphalt while producing transportation fuel.

In the past thirty years, innovation in new products and new processing

approaches has slowed dowoi. Even the capacity expansion has slowed down in the

United States. No grass-root refinery has been buih after 1980. The emphasis ofthe

refining industry has shifted to improving economic performance ofthe existing plants.

Increasing competition in the refining business and stricter environmental and safety

pressures have forced refiners to invest more money and time on process monitoring,

process control and optimization. Recent crude oil price decrease resulted in a sharp

decline in the refining margin, and consequently, decline in revenues and profits for

North American oil and gas companies (Beck, 1999). In order to survive in such

Page 18: 5-95 gap chart

changing market, companies needs to lower the operating cost and increase the revenue

to remain competitive. Recent industrial practices showed that advanced process control

and optimization are the way to accomplish this.

The common goal of all refiners is to provide safe, profitable, quality product

manufacturing (Pelham and Pharris, 1996). Process control and optimization have

become indispensable tools to realize this goal. Significant advancement in

instrumentation and computers has made the implementations of process control and

optimization cheaper and more reliable.

Before we go into the discussion of optimization which is the main topic of this

dissertation, we need to introduce process control first since the benefits of optimization

can not be realized without implementation of process control. Chemical Process Control

(CPC) is concerned with operating a plant such that the product quality and production

rate specifications are met in a safe and reliable manner (Riggs, 1999). Process Control

is also necessary to reach other operation objectives such as environmental protection,

equipment protection, and profit return, etc. (Marlin, 1995).

The controllers in a refinery were mostly PID Controllers 20 years ago. Two key

technical developments occurred during the late 1970s that led to dramatic acceptance

and growth in the number of advanced control systems. The two key developments were

the Distributed Control System (DCSs) and Model Predictive Control (MPC) Technology

(Pelham 1993). The DCS combines the hardware and software needed for data

acquisition and basic control fiinctions. It is based on using a number of local control

units which have their owoi microprocessors and are connected by shared communication

lines as well as connected to operator/engineer consoles, a date acquisition system, and a

general purpose computer (Riggs, 1999). The industrial standard for MPC is dynamic

matrix control (DMC) technology, which was implemented in late 1960s. Since then,

model-based process control has become a common industrial practice in refining

industry. It has been estimated about 3%o incremental profit can be realized through

implementing model-based process control (Ellis, 1998).

Page 19: 5-95 gap chart

The basic task for a controller is to maintain a process variable at a given setpoint.

Since every refining company needs to make profit, these setpoints must be as

economically favorable as possible. This economic target can be expressed in different

form: largest production, the greatest profit, the minimum cost, and the least energy usage

(Edgar and Himmelblau, 1988).

Now the question is raised: how to choose those setpoints which make the process

most profitable while observing process constraints and meeting all product

specifications? This is an optimization problem. In some refineries, the values of those

setpoints are set by operators and process engineers based on their experience and

intuitions. Such decision making approach can not be consistent due to different

backgrounds of people making the decisions.

In a modem refinery, a more systematic approach is followed. Usually, there is

one experienced individual known as a planner/scheduler. The responsibility of this

individual is to develop an operating plan for the next several days, given current levels

of crude stocks, operating capacities, offtakes and inventory constraints (Pelham and

Pharris, 1996). He may use linear programming as a pragmatic guideline. Today, large-

scale linear programming (LP) technology is well established with respect to planing and

scheduling and no major change in that status is anticipated (Pelham, 1993). After the

solution is found using an LP, the planning engineer sends the values in the solution as

setpoints to engineers in unit operation. Usually, the planer/scheduler gives a range for

the setpoints to a process variable instead of a single value in order to enable some

flexibility. In plant operation, the engineers in a specific unit then use the information to

choose the setpoints for each respective control loop. They also need to decide the

setpoints that are not provided by the LP.

There are several obstacles in using LP for operation scheduling in a refinery: (1)

If the refinery is operated at constant conditions and LP model well represents the

refining processes at the fixed operating conditions, LP is the answer to scheduling

problem in a refinery. However, this is rarely the case. Due to continuous change ofthe

crude quality, and fluctuating market demands and prices, the operation of a refinery is

Page 20: 5-95 gap chart

always changing to adapt to such changes. Since refining processes are inherently

nonlinear, especially those processes with reaction systems, linear models will not be able

to precisely represent those processes with linear relationship, and (2) the linear models

used by the planner/scheduler are simplified models that do not include all the details

(Jones, 1999). This simplification is done to speed up LP solution finding. The linear

models are developed by linearizing a nonlinear process at certain operating conditions.

When the operating conditions changes, the linear models become invalid. Hence, the

optimization results based on such linear models are susceptible to error, and (3) the LP

does not consider the elements of time and storage. It assumes that all activities occur

simultaneously (Hartmann, 1997), which is not the case in reality, and (4) the LP is only

suited to find an optimum at the intersections of constraints. However, it is quite possible

that the optimal solution of refinery-wide operation lies inside constraints due to the

nonlinearity ofthe optimization problems.

A possible remedy to linear programming is to update the linear models

continuously based on plant data (Cutler, 1999). However, with the increasing

complexity of refineries, it is very difficult to maintain the required data consistency

(Hartmann, 1997). In addition, the continuous updating will add tremendous workload.

On the other hand, nonlinear models have been built and single-unit optimization

has been implemented for several important refining processes, such as crude unit, FCC

unit, and gasoline blending, etc. Some software vendors have provided simulation

software packages for modeling, such as AspenPlus from Aspen Technology, PR02 from

Simulation Sciences, and HYSYS from Hyprotech, etc. Some vendors even supply

optimization packages, such as real-time optimziation (RT-Opt) from Aspen Technology,

and Romeo process optimization from Simulation Sciences. It has been estimated about

3% incremental profit can be realized through implementing real-time on-line

optimization on single unit (Ellis, 1998).

However, a single-unit optimizer covers only a subset of a large problem

(Friedman, 1995). It is hard to make the optimal solutions from single-unit optimization

to be consistent with the plant-wide operation strategy. The reason is that single-unit

Page 21: 5-95 gap chart

optimization needs correct price information ofthe intermediate streams. However, such

price information is not always available from the market since some intermediate

streams are not sold on the market. Intermediate product price is a function of quality and

product rate. It also depends on tankage consideration, lifting schedule, the current

content of intermediate tanks and finally on future crude runs (Friedman, 1995). To solve

this problem, some engineers used the shadow prices ofthe refinery LP to evaluate the

prices of intermediate products. A shadow price is the change in the optimal value ofthe

objective fimction per unit change in an active constraint. However, the shadow prices of

LP are not valid when the operation shift to different product qualities or a different

schedule. The single-unit optimization based on imprecise price information can results

in conflict with plant-wide operation strategy.

In order to overcome the inherent disadvantages of LP and single-unit nonlinear

optimization, a new approach is applied to tackle the optimization problem in the

refinery. In this work, several nonlinear single-unit models are integrated into a refinery-

wide model and this model is used in nonlinear optimization to find the optimal operating

conditions for the entire refinery. The main advantage of such an approach is that it uses

detailed nonlinear models to represent nonlinear refining processes and it does not need

prices of intermediate streams.

The primary objectives of this work are to:

1. Develop nonlinear single-unit models for important processes in a fuel-oriented

refinery.

2. Integrate developed single-unit models into a overall refinery-wide model.

3. Carry out nonlinear optimization using the developed models to find the optimal

operating conditions for the whole refinery.

4. Evaluate the incremental profit of using such optimization approach by comparing

with results from normal operating conditions.

5. Compare refinery-wide nonlinear optimization with single-unit nonlinear,

optimization.

Page 22: 5-95 gap chart

Modeling every unit in detail is formidable task if not impossible. However, a

detailed plant-wide model is not necessary for optimization to capture the major part of

the benefit obtained from nonlinear optimization just like LP does not need to use

detailed models to be helpfiil in guiding the planer/scheduler. To keep this work

manageable, the following approach is applied. Detailed models for four major units

were developed. Those units include crude unit, FCC unit, catalytic reformer, gasoline

blending. Simplified models are used to represent other units in the refinery.

The previous research on refinery modeling and optimization is reviewed in

Chapter 2. The single-unit model development and benchmarking are described from

Chapter 3 to Chapter 7. Those units including crude unit, FCC unit, catalytic reformer,

gasoline blending, gas plant and alkylation unit, etc. Most ofthe models were

benchmarked against the industrial data obtained from a fuel-oriented refinery. Chapter 3

discusses the development of a simplified first principle model of a crude unit and the

model benchmarking. In Chapter 4, the rigorous steady-state model of a Model IV FCC

unit is briefly discussed. The emphasis is on the modification ofthe existing model to fit

the FCC unit in the ftael-oriented refinery. Again, the model benchmarking is discussed.

In Chapter 5, the rigorous steady-state model of a catalytic reformer unit is briefly

discussed. The emphasis is on the modification of the existing model to fit the reformer

unit in the fuel-oriented refinery. Again, the model benchmarking is discussed. Chapter 6

discusses the modeling of gas plant, alkylation unit and other auxiliary units in the

refinery. Chapter 7 discusses the development a detailed model ofthe gasoline blending

process. Chapter 8 discusses the optimization results obtained from applying successive

quadratic programming (SQP) technique to the refinery-wide model. The comparisons

among nonlinear refinery-wide optimization, normal operating conditions and nonlinear

single-unit optimization are also given. The results of this work and recommendations for

future studies are presented in Chapter 9.

Page 23: 5-95 gap chart

CHAPTER 2

LITERATURE REVIEW

This research is divided into two parts: (1) Single-unit modeling and feed

characterization; (2) Steady-state optimization. The literature review is also divided into

the same two parts.

2.1 Single-Unit Modeling and Feed Characterization

The focus ofthe present work is the modeling and optimization of a fiiel-oriented

refinery. A typical fiael-oriented refinery includes the following major units: crude unit,

FCC unit, catalytic reformer, gasoline blending, gas plant, and alkylation unit. For a

large-scale refinery processing heavy crude, a hydrocracker or a delayed coker is often

included to process the heavy end ofthe crude. The refinery that the model was

benchmarked against is a small-scale refinery processing light crude. Hence, a

hydrocracker or delayed coker was not installed. Beside these major units, there are some

auxiliary and pretreating units in the refinery: naphtha desulfurizer, diesel hydrotreater,

Debutanizer, naphtha splitter, Claus and Scott units, Merox and Amine treater, fuel gas

mix drum, and gas turbines, etc. They are also important in the refinery-wide

optimization because their capacity may become active constraints in searching the

optimal operating conditions.

Refining process modeling is a hot topic in both industry and academia. The

industry has an edge in this area because it has large amount of operation data which

accurate and detailed models can be built on. However, refining companies rarely publish

their models. Models are built in academia for control and optimization studies. The

models in the public domain usually are based on a specific set of industrial data. Hence,

they can not be used directly in this work without further benchmarking.

In U.S., 40-50 percent of gasoline comes from FCC gasoline. Therefore, FCC unit

is a critical unit when studying the economy of a fuel-oriented refinery. Ellis (1996)

developed a steady-state model of FCC unit on the basis of a Model IV FCCU Dynamic

Simulator (McFarlane et al., 1993) while using the Mobil ten-lump yield model (Jacob et

8

Page 24: 5-95 gap chart

al., 1976) to predict product yields. The ten-lump model includes eight lumps for gas oil

components, one lump for FCC gasoline and another lump for coke and light gas. The

detailed composition modeling for FCC gasoline and light gas is not available from the

ten-lump model. Empirical correlations were used in the FCC model to calculate octane

number of FCC gasoline and the light gas composition. A detailed literature survey on

FCC modeling and optimization is included in the thesis of Ellis (1996).

Catalytic reformer produces about 20-35 percent gasoline ofthe gasoline pool.

Reformate, the gasoline blending stock produced in reformers, is the major gasoline

booster in the gasoline pool. A rigorous steady-state model of catalytic reformer was

developed by Taskar (1996). The model uses 35 components, including real chemical

species and pseudo-components, and 36 reactions to model the reaction system of a

reformer. The model provides detailed information of reformate. A detailed literature

survey on reformer modeling and optimization is included in the dissertation of Taskar

(1996).

The literature review here focuses on crude unit modeling and gasoline blending

modeling. A crude unit includes pretreating facilities, preheat train, fiimace,

atmospheric tower, vacuum tower and separation train. A detailed model, which includes

all the units, has not been presented in literature. Instead, most ofthe publishing research

focused on the tray-to-tray rigorous modeling ofthe atmospheric and vacuum towers.

Edmister (1955) proposed an integral method for making petroleum distillation

calculations. Taylor and Edmister (1971) applied the method to a gasoline rerun column.

However, this method requires a curve of K versus mole fraction, which is not easily

obtained. In addition, it uses a trial and error in root searching, which causes slow

convergence in the digital computation. Cechhetti et al. (1963) studied a 62-stage Exxon

crude tower and built a steady-state model which uses 6-method of convergence. Hess et

al. (1977) and Holland and Liapis (1983) used 2N Newton-Raphson method to solve this

steady-state model. In the above studies, the pseudo-component approach is used in vapor

liquid equilibrium (VLE) calculations. The crude is divided into 35 pseudo-components

in order to represent the true-boiling-point (TBP) curve. Cechetti et al. (1963) first

Page 25: 5-95 gap chart

presented the physical properties ofthe pseudo-components. Hess (1977) curve-fitted

those physical properties into formulas to correlate the equilibrium K values and

enthalpies as a function of temperature only.

Since the conventional bubble-point algorithm in temperature calculation is

inherently unstable in the face of enthalpy inversion caused by the extremely wide-

boiling nature ofthe crude feed, Chung and Riggs (1995) proposed a dynamic stagewise

adiabatic flash (DSAF) algorithm which efficiently provides stable solutions for an

extensive operating range. A dynamic model was constructed based on DSAF algorithm

and was used for control studies.

All the above studies used sequential modular approach in modeling. Mizoguchi

et al. (1995) constructed a dynamic model using open-equation approach for control

study. The model is based on a crude unit in the Petro-Canada Ontario Refinery. It uses

34 pseudo-components to represent the crude feed. The complete model has 2297

differential/algebraic equations, 263 fixed external variables, 9 external manipulated

variables, and 2297 dependent variables.

The present work concentrates on the economic optimization ofthe whole

refinery instead of a single unit. Hence, it is not strictly necessary to build a tray-to-tray

rigorous model for atmospheric tower and vacuum tower. The most important

characteristic that a crude unit model must have is that it should be able to predict the

volumetric flow rate and properties ofthe products accurately. Watkins (1979) presented

a design method which can be used to build a first-principle steady-state model. This

approach uses material and energy balance to calculate the sfreams in and out of each

fractional section between adjacent side-draw trays. Empirical correlations are used to

evaluate the properties of feed and products. Since this approach serves the main purpose

of this work, the crude unit has been modeled using the approach outlined by Watkins

(1979).

Gasoline cuts coming from various upstream process units are blended into

various grades of gasoline in the gasoline blending unit. The purpose ofthe gasoline

blending model is to predict the property specifications for various grades of gasoline

10

Page 26: 5-95 gap chart

from the properties of incoming gasoline cuts. Some specifications ofthe final product

can simply be calculated as the summation of corresponding properties ofthe blending

agents. Such specification include oxygen content, sulfur content, benzene content, etc.

Other specifications need a more complex, usually nonlinear, model in order to make

accurate predictions. Such specifications include octane number, Reid Vapor Pressure

(RVP), and volatility. The refinery considered in this work mainly produces regular

grades of gasoline. Reformulated gasoline which is used in some areas of U.S. has more

specifications. The calculations of extra specifications of reformulated gasoline are not

included in the gasoline blending model.

There are two approaches used in octane number calculation ofthe gasoline

blending model. The first approach is to convert the octane numbers of gasoline cuts into

blending octane values and calculate the octane of final gasoline on a volumetric average

basis. This approach is widely used in industry. Gary and Handwerk (1984) and

Unzelman (1996) gave the blending octane values of some gasoline cuts. However,

blending octane numbers are usually derived from regression analysis of a small data set,

e.g., those derived from gasoline of a single refinery. Muller (1992) used the concept of

"excess octane number" and formulated the equations to calculate the deviation of

blending octane number from ideal blending. The excess octane numbers are derived

directly from original refinery blending data. This approach is very similar to the

blending value approach. Zahed et al. (1993) presented ein empirical correlation of

calculating gasoline octane number using constants regressed from a specific refinery.

The disadvantage ofthe blending value approach is that blending values can only

be to calculate the blends where the blending values are derived from. The blending value

of one gasoline cut may change dramatically when blending is changed. The advantage of

the approach is that if the blending values are derived from a complete gasoline blending

study in a refinery, the blending value approach is more accurate for that specific refinery

than general methods introduced below.

A more theoretical approach is the interaction method. It is well known that the

octane numbers do not blend linearly due to nonlinear interactions among gasoline cuts.

II

Page 27: 5-95 gap chart

If the interaction coefficients among various gasoline cuts can be calculated, the octane

number of blended gasoline can be predicted accurately. Schoen and Mrstik (1955) first

presented a graphical correlation for predicting octane numbers of binary blends from the

octane numbers and olefin content ofthe two base blending stocks. Stewart (1959) used

the same approach but made the method more self-sustained and expanded it to multi-cut

blends. The interaction method is first proposed by Morris (1986). The key to the

accuracy ofthe interaction method is having accurate interaction coefficients. Interaction

coefficients can be derived from an octane blending study wherein accurate octane data

are obtained on all components and all 50:50 blends (Morris, 1986). Morris et al. (1994)

then expanded the interaction method to multi-agents blending.

The disadvantage of Morris's approach is that interaction coefficients can only be

used for the blending ofthe specific set of gasoline cuts where the interaction coefficients

were derived from. The interaction parameters not only depend on types ofthe gasoline

cuts, but also depend on the octane levels and octane difference ofthe gasoline cuts in the

blending. The value of interaction coefficient between two gasoline cuts can vary from

large negative to large positive in different blending situations (Twu and Coon, 1996).

The ideal solution is to have an accurate generalized method to predict interaction

coefficients without the need for a blending study. Such a method is commercially

available but is not available in open literature (Morris, 1985).

The interaction method proposed by Morris only considers the interaction

coefficients among blending cuts. It does not use the information ofthe composition of

gasoline cuts. Twu and Coon (1996) proposed a component-oriented interaction approach

that is general and can be used in any gasoline blending without blending studies. This

approach only needs the octane numbers of gasoline agents and the information ofthe

concentrations of olefins, aromatics, and saturates in the gasoline cuts. A universal set of

the binary interaction parameters is given. Twu and Coon (1997) further extend this

method to be more consistent in methodology by applying the same binary interaction

parameters to components in each gasoline cut and their blends. Due to the lack of a

12

Page 28: 5-95 gap chart

complete blending study ofthe refinery considered in this work, Twu and Coon's

approach was used in the octane number calculation in this work.

Reid vapor pressure (RVP) is widely used as a criterion to measure the volatility

of gasoline. Thus, the accurate prediction of RVP is critical in product blending and

refinery-wide optimization. A similar blending value approach is widely used in industry.

Chevron Research Company developed an empirical method which uses vapor pressure

blending indices (VPBI), which is a fiinction ofthe RVP ofthe individual blending cuts.

The table of VPBI versus RVP is given in Gary and Handwerk (1984).

Stewart (1959) formulated the first theoretical method to predict RVPs of ftiel

blends. It is based on the standard RVP test defined by the American Society for Testing

and Materials (ASTM) under the designation ASTM D323-56. It makes a VLE

calculation with respect to the Reid test conditions with the assumption of ideal gas phase

and uses the Sounders' state of equation to calculate the vapor-liquid equilibrium K

values. However, this method ignores the presence of air and water vapor in the Reid

test. It also assumes that the volatile components have the molar density of butanes and

the nonvolatile components have the thermal expansion characteristics of n-octane.

Vazquez-Esparragoza et al. (1992) followed the same approach but used the more

accurate Soave-Redlich-Kwong equation of state in VLE calculation. Stewart's approach

represents gasoline blending cuts in pseudo-components and requires the physical

properties of those pseudo-components, which can only be obtained through empirical

correlations. This adds to the inaccuracy of this approach. Since Chevron's VPBI method

is accurate enough (Gary and Handwerk, 1984) and widely used in the refining industry,

the VPBI method is used in the RVP calculation in the present work.

Empirical correlations are used in the feed characterization for the FCC unit and

the reformer unit. The most important feed information is the compositions of paraffins,

olefins, naphthenes, and aromatics (PONA). The PONA information in FCC feed can be

calculated by TOTAL method (Dhulesia, 1986) and n-d-M method (ASTM, 1985). The

reformer feed characterization uses the crude assays obtained from the refinery

considered in this work.

13

Page 29: 5-95 gap chart

After a refinery-wide model had been developed, optimization studies were

carried out to find the optimal operating conditions for the whole refinery.

2.2 Refinery-Wide Optimization

Refinery-wide optimization involves many decision variables and constraints.

Linear Programming (LP) is the most widely used technique in refinery operation

optimization, which is called planning and scheduling in industry. The linear

programming was first proposed by Dantzig in 1947 to refer to the optimization problems

in which both the objective function and the constraints are linear (Dantzig, 1963). LP

problems exhibit the special characteristic that the optimal solution ofthe problem must

lie on some constraints or at the intersection of several constraints. Dantzig first proposed

the most popular algorithm in LP called the Simplex algorithm in 1947. Symonds (1956)

used an LP to solve a simplified gasoline refining and blending problem. The advantage

of LP is its quick convergence and ease to implement. However, it is suitable only for

linear and nearly linear process.

One method to solve a nonlinear problem by using LP is to repeatedly linearize

the objective function and constraints at some estimate ofthe solution (Edgar and

Himmelblau, 1988). This procedure is the basis of a method called successive linear

programming (SLP). According to Edgar and Himmelblau (1988), SLP has been

demonstrated to be successful on problems with a moderate degree of nonlinearity. The

disadvantage of SLP is that it may converge very slowly when the optimum lies in the

interior ofthe feasible region and when there are a large number of nonlinear variables.

Generalized Reduced Gradient (GRG) algorithm is a popular nonlinear

optimization method used in industry. Basically, GRG uses a reduced form of gradient to

find the search direction. It defines new variables that are normal to the constraints and

expresses the gradient in terms of this normal basis. GRG may sometimes have difficult

to return from an infeasible point to the feasible region during the search. Generally, the

performance ofthe GRG algorithm is comparable to that of Successive Quadratic

Programming discussed below.

14

Page 30: 5-95 gap chart

Successive Quadratic Programming (SQP) is widely used to solve large-scale

nonlinear problems. Edgar and Himmelblau (1988) made the statement that SQP might

be the best in solving nonlinear programming problems. Riggs (1994) suggested using

SQP for large-scale problems. SQP is the method used in the leading on-line optimization

software package, RT-OPT, developed by Aspen Technology, Inc. SQP approximates

the objective function locally by a quadratic function and linearized constraints. The

search direction is decided by solving the quadratic programming subproblem. After that,

a minimization algorithm is used to calculate a step size in the search direction. Several

tries of line search are employed in deciding the appropriate step size. SQP repeatedly

applies the procedure until it successfully finds a point that satisfies the first-order Kuhn-

Tucker condition or fails after certain number of tries. Detailed mathematical proof on

SQP can be found in Gill et al. (1981). SQP usually requires less iterations and converges

much faster to interior optimum points than Successive Linear Programming. It should be

noted that SQP may go through an infeasible path while searching for the optimum. The

bounds and linear constraints are always satisfied along the search while the nonlinear

constraints will not generally be satisfied until an optimal point is reached (Gill et al.,

1986). The NPSOL 4.0 (Gill et al., 1986) software package is used as the optimization

engine in this work.

15

Page 31: 5-95 gap chart

CHAPTER 3

CRUDE UNIT MODELING

3.1 Introduction

The crude unit separates the crude into various product cuts with different boiling

ranges. The downstream units, such as FCC unit, reformer, further process these cuts to

make final products. A block diagram of a fuel-oriented refinery is shown in Figure 3.1.

It can be seen that downstream units receive their feed from the crude unit.

The crude unit is the first major processing unit in almost all the refineries. The

products from crude oil are usually characterized by their boiling range. Many properties

of crude cuts are related to the boiling range. For example, molecular weight and sulfur

content increase with boiling range. The aromatics content also increases with boiling

range while the pEiraffin content decreases with boiling range.

Distillation is the cheapest and the easiest method to separate crude into different

cuts. Higher efficiencies and lower costs are achieved if the crude oil separation is

accomplished in two steps: first, by fractionating the total crude oil at a atmospheric

pressure; then, by feeding the high-boiling bottoms fraction from the atmospheric still

(atmospheric tower) to a second fractionator (vacuum tower) operated at a high vacuum

(Gary and Handwerk, 1984). The flow diagram ofthe atmospheric tower and the vacuum

tower is shown in Figure 3.2.

From the point of view of plant-wide economy, the cut points between adjacent

cuts in the distillation columns ofthe crude unit are among the most important operating

decision points. If the crude cuts are not specified properly, the downstream processing

units may have trouble in processing the cuts to make products within the specifications.

The quality of crude cuts may also affect the normal operation of downstream units, such

as catalyst poisoning, equipment corrosion, etc. Therefore, operating atmospheric tower

and vacuum tower properly is crucial for the entire refinery operation.

16

Page 32: 5-95 gap chart

u

0)

-o c

17

Page 33: 5-95 gap chart

^

Atmospheric Tower

^

29

- • Vapor Distillate

f ^ Water Liquid * • Light Naphtha

U^ 4 Steam

~:©*~fieavy Naphtha

Steam Kerosene

— •

Steam

-:0^ Diesel

Steam

— * ^ ^ — • Atmospheric Gas Oil

*• To jets

>• Light Vac. Gas Oil

*• Heavy Vac. Gas Oil

Overflash

*• Vac. Bottoms

Figure 3.2 Schematic of a Fuel Type Crude Unit.

18

Page 34: 5-95 gap chart

The crude unit modeling in this work focuses on the atmospheric tower and the

vacuum tower, the core ofthe crude unit. The modeling procedure as described in

Watkins (1979) and Lin (1988) has been followed in most ofthe work.

In the crude unit, the crude from crude storage tanks first goes through

dewatering, desalting and then exchanges heat with the side-stream products and

pimiparounds from the atmospheric tower and the vacuum tower. The crude is further

heated in a furnace before entering the bottom ofthe atmospheric tower called flash zone.

The crude then partially vaporizes in the flash zone due to the pressure drop at the

injection point.

The atmospheric tower has only a rectifying section. The vapor formed in the

flash zone flows upward and is separated into overhead gas, light naphtha (LN), and

several side-draw products: heavy naphtha (HN), light distillate (LD), heavy distillate

(HD), and atmospheric gas oil (AGO). These side-draw products are drawn out ofthe

tower from the side-draw trays ofthe atmospheric tower. The light naphtha is further

separated into liquidified petroleum gas (LPG) and light straight-run gasoline (LSR). The

heavy naphtha is fed to the reformer to make reformate, a gasoline blending stock. The

light distillate is further treated to make jet fuel. The heavy distillate goes through the

hydrodesulfurization unit to make diesel. The AGO goes through the rerun unit to recover

some diesel components and the rest ofthe AGO is fed to the FCC unit to produce FCC

gasoline.

The nonvolatile part ofthe crude, which is called the reduced crude, exits from

the bottom ofthe atmospheric tower and is fed to the vacuum tower for ftirther

separation. The function of vacuum tower is to maximize the extraction of distillate

liquids from the raw crude feed. The reduced crude is first heated up in a vacuum tower

furnace and then enters the flash zone ofthe vacuum tower at the bottom. The reduced

crude vaporizes in the flash zone due to pressure drop. The vapor flows upward and is

separated into light vacuum gas oil (LVGO) and heavy vacuum gas oil (HVGO) and

product cuts are drawn from the tower on side-draw trays. The nonvolatile part ofthe

vacuum tower feed leaves the bottom ofthe vacuum tower and goes for further treatment.

19

Page 35: 5-95 gap chart

The LVGO and the HVGO are combined and fed to FCC unit. The vacuum residue goes

through the residue oil solvent extraction (ROSE) unit to recover the gas oil in the

vacuum residue. The rest of residue is further treated to make No. 6 fuel oil.

The crude unit model in this work includes an atmospheric tower, a vacuum

tower, a debutanizer, a rerun unit, a ROSE unit and a naphtha splitter. The inputs to the

crude unit model are crude assays and crude feed rate, the ASTM 95% points ofthe light

naphtha, the heavy naphtha, the light distillate and the heavy distillate, the true boiling

point (TBP) cut point ofthe heavy vacuum gas oil, atmospheric tower fiamace outlet

temperature, and vacuum tower furnace outlet temperature. The model calculates the

volume and the properties of those side-draw products and the vacuum bottom stream. It

also calculates the temperature profile in the towers.

Two standard tests are defined here since they will be often referred to in the

follow text. The ASTM 95%) point is the temperature corresponding to the 95 vol.%) on

the distillation curve measured by the distillation test defined by the American Society

for Testing and Materials (ASTM) under the designation of ASTM D86. The ASTM

method D86 is an easy test that is run in an Engler-flask at atmospheric pressure with no

trays or reflux between the stillpot and the condenser. TBP curve is measured by the

ASTM D2892 test, which uses a distillation column with 14 or more theoretical stages

and reflux ratio of 5 or more. Temperature at any point on the TBP-volumetric yield

curve represents the actual boiling point ofthe hydrocarbon material present at the

volume percentage corresponding to that point. ASTM D2892 test is more expensive than

ASTM D86 test.

3. 2 Atmospheric Tower Modeling

The schematic ofthe atmospheric tower model is shown in Figure 3.3. The text

box with gray background has internal iterations inside. The text box with white

background does not have internal iterations. The indices ofthe text boxes with gray

background are referred to when discussing some calculations with iterations in the

following text.

20

Page 36: 5-95 gap chart

Start

Side-draw Product Specifications,

ASTM 95% point TBP 100% point

Calculate the mixed crude's curves of TBP, API gravity,

and sulfur content

Operating pressure, Crude Temperature,

steam rate

Crude Assays, Crude Makeup

Calculate the properties of side-draw prodcuts

Steam rate, side stripper type

Calculate flash zone temperature, Calculate the

extra energy left in the tower after flash zone

(1)

Calculate the properties of unstripped product

witlidrawn from the draw trays (2)

Draw tray location Calculate the temperature of

the draw trays (3)

Calculate the temperature of the condenser

Calculate the temperature of side-draw products

(4)

End

Figure 3.3 Schematic ofthe Model ofthe Atmospheric Tower ofthe Crude Unit.

21

Page 37: 5-95 gap chart

3. 2.1 Feed Characterization

The refinery that the model was benchmarked against processes four different

crude oils: Crude A, Crdue B, Crdue C, Crdue D. They are from different resources with

slight different properties. The refinery uses a Linear Programming (LP) technique for

scheduling and planning. In the present work, nonlinear optimization methodology is

used and results are compared with LP report. The normal crude unit feedstock makeups

are given in Table 3.1.

Table 3.1 Feedstock Makeups.

Crude Type

Crude A

CrudeB

CrudeC

CrudeD

API Gravity

37.4

34.2

33.7

36.1

Sulftir,

0.08

0.54

1.04

0.32

Wt. Summer Mode Vol.%

17

8

48

27

Winter Mode Vol.%

0

0

0

100

All four crudes processed in the refinery can be characterized as light crude type.

The specific gravity of these crudes range from 33.7 to 37.4. Among these crudes, only

crude C is a sour crude and has a sulfur content of 1.04 wt.%. Other crudes have very low

sulfur content, less than 0.54wt.%.. The refinery was designed to process light crude with

low sulfur content. It can not process heavy crudes or crudes with high sulfur content.

The volumetric flow rate of a side-draw product is equal to the volumetric flow

rate ofthe total crude feed multiplied by the volumetric percentage ofthe side-draw

product in the crude feed. Once the initial boiling point (IBP) and end boiling point

(EBP) ofthe side-draw product are known, the volume percentage ofthe side-draw

product can be found from the true boiling point (TBP)-volumetric yield curve ofthe

whole crude as shown in Figure 3.4.

22

Page 38: 5-95 gap chart

1000

^ 800 £ 3

5 600 (U Q. E jfl) 4 0 0

200

LN

HN LD

HD

AGOLVGO H V G p / Resid.

10 20 30 40 50 60 volume%

70 80 90 100

Figure 3.4 Crude Cuts on a TBP-Volumetric Curve.

The volume percentage of a side-draw product is calculated using the formula

given below:

%V,=%V,,,,-%V^j,„ (3.1)

where

i - side-draw product; light naphtha (LN), heavy naphtha (HN), light distillate (LD),

heavy distillate (HD), atmospheric gas oil (AGO), light vacuum gas oil (LVGO), heavy

vacuum gas oil (HVGO),

%)Vi- volume percentage ofthe side-draw product i in the crude,

%oVijBp, %iVi,EBP- volume percentage corresponding to the IBP ofthe side-product i on

the TBP-volumetric yield curve ofthe crude, respectively.

Hence, the TBP-volumetric yield curve ofthe crude feed is essential to calculate

the volumetric flow rates ofthe side-draw products.

The crude assays of these four types of crude oils were obtained from the refinery.

The TBP-volumetric yield curve ofthe crude oil entering the crude unit is calculated

23

Page 39: 5-95 gap chart

The crude assays of these four types of crude oils were obtained from the refinery.

The TBP-volumetric yield curve ofthe crude oil entering the crude unit is calculated

based on the TBP curves of each crude and its volumetric fractions in the feed. Ideal

mixing is assumed here because these crude oils themselves are complex mixtures where

deviations from the ideal mixing are assumed to be cancelled out. In addition, they are all

low sulfur, light crudes with similar properties. The equation to calculate the TBP curve

of mixed crude oil is given below:

v(T) = Y^v,(T)xvol,, (3.2)

where

i - crude type: A, B, C, and D,

v(T)- volumetric percent ofthe mixed crude with boiling temperature lower than T,

vol.%,

Vi(T)- volxmietric percent of crude i with boiling temperature lower than T , vol.%,

vol,- volume fraction ofthe crude i in the mixed crude.

The TBP curve ofthe mixed crude oil used in Summer Mode operation is shown

in Figure 3.5. For the convenience of digital computations, the TBP curve ofthe mixed

crude is regressed using a polynomial expression. The regression was carried out in

MathCad Plus 6.0 from Mathsoft Incorporation. The order of a polynomial expression is

selected by looking at the mean square errors ofthe predictions. In this work, a

polynomial expression usually has the order in the range of 5 to 7. For this polynomial

expression and any polynomial expression mentioned in the following text, the order and

the constants ofthe polynomial expression are listed in Appendix A. It should be noted

that different TBP curves are used in the Summer Mode and the Winter Mode due to the

fact that different crudes are used for different operation Mode.

24

Page 40: 5-95 gap chart

" 600

10 20 30 40 50 60

volume%

70 80 90 100

Figure 3.5 TBP Curve ofthe Crude Used in Summer Mode.

Another curve required by the crude unit model is the API-volumetric yield curve.

The relation between API Gravity and specific gravity is given below:

141 5 API = ^^!±^-131.5,

SPG (3.3)

where

API- API gravity,

SPG- specific gravity.

The API-volumetric yield curve ofthe mixed crude is calculated from the API-

volumetric yield curves ofthe crude sources. The procedure of calculating the API-

volumetric yield curve ofthe mixed crude is given below:

a. Approximate the API-volumetric yield curve of a particular crude by a polynomial

expression.

25

Page 41: 5-95 gap chart

b. Use a certain number of intermediate temperature points to represent the boiling range

ofthe crude. For each temperature, calculate the corresponding volume percentage from

the TBP-volumetric yield curve ofthe crude.

c. For each volume value, find the corresponding API value using the API-volumetric

yield curve ofthe crude.

d. Convert API to specific gravity using equation 3.3.

e. Repeat step a to step b for all four types of crudes.

f. The API ofthe mixed crude is calculated using the following equation:

SPG(v) = Y, SPG^ (v) X vo/. , (3.4)

where

SPG(v)- specific gravity of mixed crude,

SPGi(v)- specific gravity of crude i,

voli- volume percentage of crude i.

g. Convert specific gravity to API gravity.

h. Construct API-volumetric yield curve using the calculated API and approximate it by

a polynomial expression.

The API-volumetric yield curve ofthe mixed crude oil used in Summer Mode operation

is shown in Figure 3.6.

The sulfur-volumetric yield curve ofthe mixed crude is also calculated using

similar procedure. It should be noted that the sulfur concentration is expressed as weight

percentage instead of volume percentage. The sulfur-volumetric yield curve ofthe mixed

crude oil used in the Summer Mode is shown in Figure 3.7. The sulfur-volumetric yield

curve was also regressed using a polynomial expression. The polynomial expressions

developed above are then used repeatedly in the atmospheric tower model and vacuum

tower model.

The crude processed in Winter mode operation has one crude type and no mixing

calculation is required.

26

Page 42: 5-95 gap chart

0 10 20 30 40 50 60 70 80 90 volume%

100

Figure 3.6 API Curve of the Crude Feed Used in Summer Mode.

2.5 T

Figure 3.7 Sulfur Curve ofthe Crude Feed Used in Summer Mode.

27

Page 43: 5-95 gap chart

3.2.2 Side-Draw Product Calculation

The boiling range specification for a side-draw product is usually defined by

ASTM 95% point. Since TBP test is more expensive, TBP test is only run on crude oils

but not on product fractions. Hence, the side-draw product specification is defined using

ASTM data instead TBP data. Conversion from ASTM 95% points to TBP cut points in

the model is required in order to use the TBP-volumetric curve ofthe crude feed.

It is assumed that the atmospheric tower and the vacuum tower are able to

separate the side-draw products according to the ASTM 95%) specifications. This

assumption is made based on the fact that the values ofthe ASTM 95% points are limited

in a narrow range in the optimization studies and the crude unit has enough separation

capacity to satisfy the separation requirements in this range. This assumption is

confirmed by the operating personel in the refinery considered in this work. The

properties ofthe side-draw products can then be calculated using the ASTM 95% point

specifications without considering whether these specifications can be reached or not.

Since the volume percentages ofthe side-draw products in the crude feed can only

be obtained from the TBP ciuve ofthe crude feed, it is necessary to convert the

specifications in the form of ASTM 95% points to TBP end points using the relation

curve given by Watkins (1979). The curve has been regressed into a polynomial

expression given below:

TBP end po int = X ^' ' i^^^TM 95% po int) , (3.5) i=0

where

Ci- constants ofthe polynomial expression.

In order to decide the volume percentage of a side-draw product in the crude feed,

the initial boiling point (IBP) is also required. It should be noted that the overlap ofthe

TBP curves of two adjacent side-draw products can not be avoided since two side-draw

streams can not be completely separated in the atmospheric tower and the vacuum tower.

28

Page 44: 5-95 gap chart

In other words, there are always some components in the boiling range ofthe lighter

stream entering the heavy streams and vice versa. Therefore, the IBP ofthe heavier

product is always lower than the end boiling point (EBP) ofthe adjacent lighter product.

The extent ofthe overlap represents the separation efficiency in the atmospheric tower.

The smaller the overlap, the better the separation. The separation is defined by Gap (5-

95) ASTM, which is the temperature difference between the ASTM 5% point ofthe

heavier product and the ASTM 95% point ofthe adjacent lighter product. The values of

Gap (5-95) ASTM in normal industrial practice (Watkins, 1979) are used here for all

pairs of adjacent side-products except the pair ofthe heavy distillate and the atmospheric

gas oil. The value from the Lin (1988) is used for the Gap (5-95) ASTM between the

heavy distillate and the atmospheric gas oil uses. The Gap (5-95) ASTM for each pair of

adjacent side-draw products is listed in Table 3.2.

Table 3.2 Gap (5-95) ASTM Temperature between Adjacent Side-Draw Products in the Atmospheric Tower.

Stream Gap Spec, (5-95) ASTM, °F

Light Naphtha- 25

Heavy Naphtha

Heavy Naphtha- 35

Light Distillate

Light Distillate- 10

Heavy Distillate

Heavy distillate- 5

Atmospheric gas oil

The Gap (0-100) TBP, which is the temperature difference between the 0% TBP

point ofthe heavier product and the 100%) TBP point ofthe adjacent lighter product, can

be calculated from corresponding Gap (5-95) ASTM temperature using the conversion

curves given by Watkins (1979). The curve has been regressed into a polynomial

expression shown below:

29

Page 45: 5-95 gap chart

Gap {O -1 GO) TBP = X z • [Gap (5 - 95) ASTM] , (3.6) ; • = «

where

Cj- constant ofthe polynomial expression.

Once the values ofthe Gap (0-100) TBP are knowoi, the IBP of each side-draw

product is calculated from the EBP ofthe adjacent lighter product using the following

formula:

IBP of heavier product = EBP of adjacent light product - Gap (0-100) TBP, (3.7)

The IBPs of all side-draw products are calculated using equation 3.7 except the light

naphtha, the lightest stream. The IBP ofthe light naphtha is the IBP ofthe whole crude

oil.

The TBP cut point of a side-draw product is then calculated using following

equation:

TBP cut point = -(t,„„, +to„), (3.8)

where

tiooL- TBP 100% point ofthe lighter stream, ""F,

toH- TBP 0% point ofthe heavy stream, ""F.

This methodology has been established by analysis of operating data which has shown

that, for well stripped sidestreams, the volume interchange between two streams around

the TBP cut point are equal (Watkins, 1979). The TBP cut points can then be applied to

determine the volume percentage of each side-draw product in the crude feed from the

whole crude TBP-volumetric curve.

30

Page 46: 5-95 gap chart

The volumetric flow rates of all side-draw products except the AGO can be

calculated by multiplying the volume percentage of the side-draw product by the

volumetric flow rate ofthe crude feed. For the AGO, the ASTM 95% point, which

determines the total volume of products from the atmospheric tower, is not an

independent variable once the atmospheric furnace outlet temperature is defined. This is

because that the total volume ofthe top product and the side-draw products is

approximately equal to the volume ofthe feed vaporized in the flash zone, which is

determined by the temperature and pressure in the flash zone. The total volume ofthe

product can then decide the ASTM 95% ofthe AGO. Hence, the ASTM 95% point ofthe

AGO is not an independent variable in the atmospheric tower model. Instead, it is

calculated in the crude unit model according to the conditions in the flash zone ofthe

atmospheric tower. At the beginning ofthe crude unit model calculation, an initial value

ofthe ASTM 95%) point ofthe AGO is set and the final value is calculated in the model.

The mass flow rate ofthe side-draw products can be obtained from the

corresponding volumetric flow rates and API gravity values. The API gravity for a crude

cut with narrow boiling range is calculated by applying the middle-point temperature to

the API-volumetric curve. Since the API-volumetric curve ofthe whole crude is very

close to a slight line over a narrow temperature interval, this approximation gives almost

the same value as the more accurate approach in which numerical integration is used

leading to a high computational load. However, for a wide cut, the API-volumetric curve

is not a straight line over the boiling range ofthe cut. Hence, the weight ofthe cut is first

calculated by integration along the volumetric-API curve over its boiling range. The TBP

boiling range ofthe side-draw product is divided into 10 intervals. The API gravity value

of each interval is obtained from the API-volumetric curve and is converted to specific

gravity. Simpson's rule is then used to calculate the weight ofthe side-draw product by

integrating along volumetric-API curve over its boiling range. The API gravity ofthe cut

is then determined by the volume and the weight ofthe cut. The weights of all side-draw

products are obtained from the API-volumetric curve using either ofthe two methods

introduced above.

31

Page 47: 5-95 gap chart

The TBP curve of each side-draw product is the same as the portion ofthe whole

crude TBP curve between the IBP and EBP ofthe side-draw product. For example, the

TBP curve ofthe heavy naphtha is the same as the portion ofthe whole crude TBP curve

between the IBP and EBP ofthe heavy naphtha.

The ASTM curve of a side-draw product is an important property. The ASTM

curve can be used to calculate other properties ofthe side-draw product. The ASTM

curve of each crude cut is converted from the TBP curve using the correlation given by

Daubert (1994). Daubert gave the procedure to convert an ASTM curve to a TBP curve

explicitly. Converting a TBP curve to an ASTM curve is done by reversing the

procedure. The conversion from a TBP curve to an ASTM curve follows the steps above:

a. Calculate the TBP 50% point from ASTM 50% point using the following formula:

/ ASTM(50) = (TBP( 50 ))""'', (3.9)

0.87180

where

TBP(50)- TBP distillation temperature at 50 vol.% distilled, °F,

ASTM(50)- ASTM distillation temperature at 50 vol.% distilled, °F.

It is claimed that equation 3.9 best correlates data with ASTM 50% point temperatures

below 480 °F.

b. Calculate the temperature difference among the points on the ASTM curve from the

temperature difference among the points on the TBP curve using the formula given

below:

(AASTM)^ = -(ATBP)i B, (3.10) A

where

(A TBP)i- difference in TBP distillation temperatures between two cut points, °F,

32

Page 48: 5-95 gap chart

(A ASTM)i- difference in ASTM D86 distillation temperatures between two cut points,

°F,

The values of A and B are given in Table 3.3 for various cut point ranges with the

maximum values of (A ASTM)i that can be applied in the correlation.

Table 3.3 A and B Variables for Equation 3.10.

Number

1

2

3

4

5

6

Cut-point range

100% to 90%

90% to 70%

70% to 50%

50% to 30%

30% to 10%

10%toO%

A

0.11798

3.0419

2.5282

3.0305

4.9004

7.4012

B

1.6606

0.75497

0.82002

0.80076

0.71644

0.60244

Maximum applicable (A ASTM)i, °F

-

100

150

250

250

100

c. The TBP temperatures are calculated using equations from 3.11 a to 3.11 f.

ASTM(O) = ASTM(50)-(AASTM), -(AASTM), -(AASTM),, (3.11a)

ASTM(10) = ASTM(50)-(AASTM), -(AASTM),, (3.11b)

ASTM(30) = ASTM(50)-(MSTM),, (3.11c)

ASTM{1Q) = ASTM {SO) + (AASTM)^, (3.1 Id)

ASTM(90) = ASTM (50) + (AASTM), + (MSTM),, (3.1 le)

ASTM(IOO) = ASTM(50) + (AASTM), + (AASTM)^ + (AASTM),. (3.1 If)

It is claimed that this method is more accurate than other methods (Daubert, 1994).

The volume average boiling point (VABP) and mean average boiling point

(MeABP) are calculated from the ASTM curve using the correlation given by Lin (1988).

The VABP is calculated from the following formula:

33

Page 49: 5-95 gap chart

-., „„ ASTM(IO) + ASTM(30) + ASTM(50) + ASTM(70) + ASTM(90) VABP =

(3.12)

MeABP can not be calculated by its definition. It is calculated from VABP using

the following formula:

MeABP = VABP-AT, (3.13)

where

AT- temperature difference between VABP and MeABP, " C.

AT is a function ofthe VABP and the slope ofthe ASTM curve S. The slope ofthe

ASTM curve S is calculated using the following formula:

^_ASTM(90)-ASTM(10)

90-10

Where

S- Slope of ASTM cruve, °C/%.

AT is calculated using following formula:

lnAT = -1.53181-0.0128-(VABP)"''"' +3.6474-S""'\ (3.15)

The units of VABP, MeABP, and AT are all °C.

The molecular weight of each side-draw product is calculated from the MeABP

and the API gravity ofthe side-draw product using the correlation given by Riazi and

Daubert (1980) using the following formula:

34

Page 50: 5-95 gap chart

Molecular Weight = 4.5673E - 5 • (MeABP)'"" • SPG-""'', (3.16)

where

SPG- specific gravity of the crude cut.

However, it is found that this formula predicts molecular weights of heavy cuts

lower than the industrial data obtained from the refinery considered in this work. Hence,

the molecular weights of crude cuts heavy than light naphtha are calculated from the

curves given by Maxwell (1950) which correlate the molecular weight of crude cut to its

MeABP and characterization factor. The characterization factor is calculated using the

formula given below:

Characterization Factor = , (3.17) SPG

where MeABP has units of °R.

Maxwell (1950) gave three molecular weight curves of crude cuts corresponding

to three regions ofthe characterization factors, 12.1-12.6, 11.7-12.0, and 11.3-11.6. The

three curves were regressed using polynomial expressions in this work. The polynomial

expression has the following formula:

n

Molecular Weight = J ] ^/" {MeABP) , (3.18)

where MeABP has units of °F. The molecular weight ofthe crude cut whose

characteristic factor falls in one ofthe three regions is calculated from equation 3.18

using the polynomial constants corresponding to it. For example, the characteristic factor

of a crude cut is 11.8, which is in the region of 11.7-12.0. The molecular weight of this

crude cut is calculated using the polynomial constants corresponding to the region of

35

Page 51: 5-95 gap chart

11.7-12.0. For all the crude cuts in this work, the characteristic factors are between 11.3

and 12.6. Hence, Maxwell method can be applied to any crude cut in this work.

3.2.3 Atmospheric Tower Furnace Calculation

The temperature ofthe crude before entering the atmospheric furnace is set at 481

°F according to the operation data obtained from the refinery considered in this work.

The heat duty ofthe fiimace is calculated as the enthalpy difference between the crude

entering and exiting the furnace. The enthalpy ofthe crude cut is calculated from the

graph of enthalpy of petroleum fractions in Watkins (1979). The enthalpy is a function of

API gravity, temperature and phase. For a particular API gravity and phase, an enthalpy-

temperature curve is given. For the convenience of digital computation, polynomial

expressions were used to represent the curves in the graph. The vaporization enthalpy of

a crude cut is calculated by evaluating the difference between enthalpies ofthe crude cut

in liquid phase and in vapor phase.

3.2.4 Flash Zone Calculation

The temperature and the pressure in the flash zone are required to calculate the

volume of total distillate products. The flash zone pressure is often set at the minimum

possible level in order to maximize the crude vaporization at a certain temperature. The

pressure in the condenser ofthe atmospheric tower and pressure drops across the

condenser and associated piping, trays, and transfer line between the atmospheric furnace

and flash zone in the atmospheric tower considered here are listed in Table 3.4.

The pressure in the flash zone is then calculated as below:

P - P + AP + AP Ci 1Q"* flashzom condenser piping lower' \-'• ^ ^)

36

Page 52: 5-95 gap chart

Table 3.4 Pressure in the Atmospheric Tower.

Condenser Pressure

Pressure drop across the

condenser and associating line

Pressure drop between the

column top and the flash zone

Pressure drop across the transfer

line

Pressure value

14.5

1.9

1.9

5

Pressure unit

Pisa

psi

psi

psi

It is assumed that in the overflash, the extra vaporization which provides reflux in

the section between the flash zone and the first side-stream product draw tray, is 2 vol.%

ofthe crude feed based on common industrial practice (Watkins, 1979). Further, it is

assumed that the bottom stream is steam stripped at a rate of 10 pound steam per barrel of

the reduced crude according to common industry practice. The overflash and steam rate

used in the refinery varies in the operation. The assumed values are in the normal

operation range ofthe refinery.

The flash zone ofthe atmospheric tower is shown in Figure 3.8.

The steps involved in the heat and material balance calculations are listed below:

a. Calculate the volume, weight and API gravity of each side-draw product except AGO

using the given ASTM 95% points and whole crude TBP-volumetric curve and API-

volumetric curve.

b. Assume an initial value for ASTM 95% cut point of AGO.

1. Calculate the volume, weight and API gravity of AGO.

2. Calculate the volume and weight of total distillate product ofthe atmospheric tower.

3. Calculate the volume and weight ofthe reduced crude.

c. Assume an initial value for the flash zone temperature. Calculate the feed flash and

bottom stripping required to yield the required volume of vapor. The curves given by

Watkins (1979) are used to calculate the volume percentage of nonvaporized crude which

37

Page 53: 5-95 gap chart

is stripped out. The curves in Watkins (1979) were regressed using polynomial

expressions.

1. Calculate the volume percentage of crude distilled in the flash zone. Equilibrium

flash vaporization (EFV) curves are constructed from the flash data calculated in

ChemCad 4.0 software package from Chemstations, Incorporation. The constants of

the EFV curves are listed in Appendix A. Flash curves between 26 psia and 48 psia

are constructed in this work. The pressure in the flash zone is then calculated by

interpolating among these EFV curves according to the volume percentage distilled.

Calculate the total enthalpy in the flash zone.

2. Calculate the total enthalpy at the furnace outlet using the given fiimace outlet

temperature.

d. Compare the total enthalpy in the flash zone and the ftimace outlet. If they do not

match, calculate the flash zone inlet temperature using secant method (Riggs, 1994).

Return to step c.

e. Compare the flash zone pressure obtained above and the flash zone pressure

calculated from equation 3.19. If they do not match, calculate new value of ASTM

95% point of AGO using secant method. Return to step b.

The flowchart ofthe calculation ofthe flash zone pressure and temperature is

shown in Figure 3.9. The above iterative procedure is inside the gray block no.I in the

schematic ofthe atmospheric tower model shown in Figure 3.3.

38

Page 54: 5-95 gap chart

Crude

fo

"fo

Furnace

Stripping Steam

Tray 5

Vapor

Liquid

Tray 4

Tray 1 Stripout

v.

Figure 3.8 Flash Zone ofthe Atmospheric Tower.

39

Page 55: 5-95 gap chart

Start

Crude Assay

Initial Guess of ASTM 95% Point of

AGO

Property Calculation of Side-draw Products except

AGO

Initial Guess of Flash Zone Temperature

ASTM 95% Points of Side-draw

Products

AGO Calculation Total Product Calculation

Reduced Crude Calculation

Flash Calculation at Furnace Outlet and Flash Zone Calculate New Flash

Zone Temperature

Yes

Yes

Calculate New ASTM 95% of AGO

No

End

Figure 3.9 Flowchart for the Calculation of Flash Zone Pressure and Temperature.

40

Page 56: 5-95 gap chart

In step c, the curves in Watkins (1979) to calculate the amount of stripout from

the steam rate are regressed using polynomial expressions given by the following

formula:

n

Stripout = ^ c, • {steam rate) (3.20) i=0

where stripout has the unit vol.% and steam rate has the units of pounds of

steam/barrels of stripped liquid.

The heat balance ofthe flash zone is then calculated using the flash zone

temperature obtained above. The vapor in the flash zone is the summation ofthe feed

vaporized in the flash zone and the steam stripout ofthe nonvaporized feed dropping

from the flash zone to the tower bottom. Thus, the enthalpy of vapor in the flash zone is

the summation ofthe enthalpies of these two streams which have been calculated above.

The temperature ofthe flash zone vapor is calculated from the enthalpy ofthe flash zone

vapor.

There are usually four trays between the flash zone and the tower bottom. The

temperature ofthe reduced crude is calculated on the basis ofthe energy balance for the

tower bottoms. The calculation is carried out using following steps:

a. Assume the temperature ofthe tower bottom is about 30°F cooler than the tower inlet

temperature. Calculate the temperature of tray 4 assuming linear temperature

distribution in the four trays between the flash zone and the tower bottom. It should

be noted that the temperature of tray 4 is different from the inlet temperature ofthe

crude feed.

b. Calculate the enthalpy ofthe stripout ofthe liquid dropping from the flash zone. The

amount ofthe stripout is calculated using equation 3.20.

c. Calculate the enthalpy ofthe overflash liquid. It is assumed that the overflash falls to

the flash zone at a temperature of 20' F cooler than the tower inlet temperature.

d. Calculate the enthalpy ofthe reduced crude by making an energy balance around the

tower bottom.

41

Page 57: 5-95 gap chart

e. Calculate the temperature drop across the four trays between the flash zone and the

tower bottom using the enthalpy ofthe reduced crude and the specific heat ofthe

reduced crude. Calculate the tower bottom temperature.

f. Compare the calculated tower bottom temperature with the initial guess. If they do

not match, return to step a with the new value ofthe tower bottom temperature equal

to the calculated tower bottom temperature.

The above iterative procedure is inside the gray block no.2 in the schematic ofthe

atmospheric tower model shown in Figure 3.3.

It is assumed that the temperature ofthe reduced crude is equal to the tower

bottom temperature. The external energy left in the tower above the flash zone is then

calculated using the equation below:

External energy = enthalpy of crude and steam - enthalpy of the reduced crude . (3.21)

The enthalpy of steam is obtained from Moran and Shapiro (1996). The data were

regressed into polynomial expression for the convenience of digital computations. The

external energy calculated here will be used in the following side-draw temperature

calculation.

3.2.5 Side Stripper Calculation

When the side-draw product liquid is withdrawn from the tower, there are some

light components in the stream which belong to lighter side-draw products. To recover

these light components, side strippers are used to strip the light components from the

unstripped product stream. There is a side stripper for each side-draw product. Although

the strippers with reboiler are sometimes used in industry, only the side strippers with

steam are used in the refinery considered in this work. Hence, only steam side strippers

were considered in the present work. The side stripper models are used to calculate the

properties of unstripped products for later use in side-draw temperature calculation. It

42

Page 58: 5-95 gap chart

should be noted that the volume and properties ofthe stripped product have already been

calculated in the procedure presented in section 3.2.2.

After an unstripped side-draw product stream is withdrawn from the draw tray, it

enters the side stripper at the top. The steam enters the side stripper at the bottom. Since

the partial pressure of crude components in the gas phase is low in the side-stripper, some

light components are vaporized and flow with the steam back to the column. The stripped

product exits from the bottom ofthe stripper.

The volume percentage of the unstripped side-draw products vaporized in the side

strippers is calculated using equation 3.20. For the unstripped product, its EBP is the EBP

ofthe stripped product while its IBP is obtained from the whole crude TBP-volumetric

curve using the temperature corresponding to the volume pecentage equal to the

summation ofthe volumetric flow rate ofthe stripout and the stripped product.

Volumetric Flow Rates of Unstripped Pr oduct = Stripout +

Volumetric Flow Rates of Unstripped Pr oduct

After the corresponding IBP and EBP ofthe unstripped side-draw product are

obtained, the API gravity ofthe unstripped side-draw product can then be obtained from

the whole crude API-volumetric curve. The mass flow rates ofthe unstripped side-draw

product are calculated using its volumetric flow rates and the API gravity.

3.2.6 Draw Tray Locations

The atmospheric tower model here represents the preflash tower and the

atmospheric tower ofthe refinery considered in this work together. Thus, the draw tray

locations ofthe atmospheric tower in the refinery are not applied here. In addition, since

the crude unit is not a common distillation column, normal shortcut methods to calculate

the number of theoretical trays are not applicable here. Actually, accurate draw tray

location is not vital in the present modeling work because material balances and energy

balances are carried out around separation sections instead of single trays.

43

Page 59: 5-95 gap chart

The number of trays for a separation section between a particular pair of adjacent

side-draw products is almost the same for every atmospheric tower. The numbers

suggested by Lin (1988) are used here. The number of trays of each separation section is

listed in Table 3.5. The total number of trays in the atmospheric tower is 38.

Table 3.5 Tray Numbers of Separation Sections in Atmospheric Tower.

Separation section

Light naphtha-heavy naphtha

Heavy naphtha-light distillate

Light distillate-heavy distillate

Heavy distillate-atmospheric gas oil

Feed-atmospheric gas oil

Bottom-Feed

Tray number

7

8

7

9

3

4

The draw tray location can then be calculated from Table 3.5. It should be noted

that the numbering is from the bottom to the top. In the crude unit ofthe refinery

considered in the work, a packing section is used between AGO side-draw tray and the

flash zone. The separation capability ofthe packing section is equal to 3-4 trays. In this

study, 3 trays are assumed. The important tray locations are listed in Table 3.6.

Table 3.6 Draw Tray Location.

Separation section Tray number

Reduced Crude Bottom

Feed tray 4

AGO draw tray 7

Diesel draw tray 16

Kerosene draw tray 23

Heavy naphtha draw tray 31

Light naphtha draw tray Top

44

Page 60: 5-95 gap chart

There are two pumparounds with this crude column. One pumparound is located

in the separation section between the heavy naphtha side-draw tray and the kerosene side-

draw tray and the other pumparound is located in the separation section between the

diesel side-draw tray and the atmospheric gas oil side-draw tray. The locations ofthe

pumparound are set according to normal industrial practice (Watkins, 1979). Usually, the

pumparound withdraws liquid from a tray above the lower draw tray, exchanges heat

with crude feed and returns the liquid to a tray ftirther up in the tower but below the upper

draw tray. In the model, the upper pumparound returns the liquid to the tray below the

corresponding upper draw tray. The lower pumparound returns the liquid to the third tray

below the corresponding draw tray.

The fimction of pumparounds is to reduce the liquid and vapor traffic at the upper

section ofthe atmospheric tower and reduce the heat load ofthe overhead condenser. The

disadvantage of using pumparounds is that several trays between the pumparound

withdrawing tray and returning tray can only be considered as only one actual tray for

fractionation purposes (Watkins, 1979). The benefit of having two pumparounds is that

two different temperature levels of heat sources are provided for crude preheating.

3.2.7 Draw Tray Temperature Calculation

The calculation procedure for draw temperature is similar for all draw trays. Only

the general methodology is described here. The sequence of draw tray calculation is from

the bottom to the top. The envelops used in the energy balance and material balance

include all the lower part ofthe tower to the bottom ofthe tray above the draw tray. A

draw tray and its envelop for energy balance and material balance are shown in Figure

3.10. The draw tray calculation follows the procedure in Watkins (1979).

The draw tray temperature is calculated using energy balance. For each energy

balance around the envelop corresponding to the draw fray, the difference between the

enthalpy ofthe vapor going up and the enthalpy ofthe liquid coming down is calculated.

This enthalpy difference is regarded as the external energy left in the tower which will be

absorbed by pumparounds above the draw tray and by overhead condenser. The external

45

Page 61: 5-95 gap chart

energy calculated in the lower draw tray can be used directly in energy balance of

the upper draw tray. This approach eliminates the need to calculate the all the inlet

streams and outlet streams below the draw tray. The energy balance for a draw tray only

requires that the inlet and outlet streams around the draw tray are calculated. Usually, the

outlet stream in the energy balance is the side-draw product stream. The inlet stream is

the steam used in the side-stripper which then enters the tower. If a pumparound exists

between the draw tray and adjacent lower draw tray, the external energy left in the tower

is subtracted by the amount ofthe heat load ofthe pumparound. The pumparound heat

loads are fixed according to common industrial practice (Watkins, 1979).

tray

Envelop containing all the lower part o f the

tower

stripout

unstriped • product

Envelope for the energy balance of a side-draw tray

Envelope containing all the lower part o f the

tower

steam

side-draw product

Figure 3.10 Energy and Material Balance Quantities at a Side-Draw Tray.

46

Page 62: 5-95 gap chart

The draw temperature calculation follows the steps below:

a. Make an initial guess for the draw tray temperature. Calculate the enthalpies ofthe

liquid and vapor leaving the draw tray. The volume ofthe vapor is the summation of

volumes of all the lighter products withdrawn above the separation section. The liquid

is the unstripped side-draw product that has already been calculated in side stripper

calculation.

b. Calculate the heat absorbed in passing across the draw tray by the liquid that is

vaporized in the side-draw product stripper. The temperature of tray above the draw

tray is set at I5°F higher than the draw tray temperature.

c. Calculate the reflux heat on the basis ofthe energy balance ofthe envelope

corresponding to the draw tray.

d. Calculate the volumetric flow rate ofthe reflux falling from the tray above the draw

tray according to reflux heat and the specific heat ofthe refiux. The specific heat of

the reflux is assumed to be the same as the specific heat capacity ofthe side-draw

product.

e. Calculate the molar flow rate ofthe reflux and all the lighter products. Calculate the

pressure ofthe draw tray according to the known flash zone pressure and assumed

constant pressure drop for each tray.

f. Calculate the hydrocarbon partial pressure according to the mole fraction ofthe

reflux, which has the same composition as the side-draw product, in the total vapor

leaving the draw tray. This approach is suggested by Lin (1988).

g. Calculate the TBP curve ofthe stripped product. Convert the TBP curve to the

Equilibrium Flash Vaporization curve (EFV) wdth the partial pressure using the

method given by Maxwell (1950). The initial point ofthe EFV curve is the bubble

point ofthe stripped side-draw product.

h. Compare the calculated bubble point with the draw tray temperature. If they do not

match, use the calculated bubble point as the draw tray temperature and return to step

a.

47

Page 63: 5-95 gap chart

The flowchart for the calculation ofthe draw tray temperature is shown in Figure

3.11. The above iterations are inside the gray block no.3 in the schematic ofthe

atmospheric tower model shown in Figure 3.3.

The temperature ofthe condenser is calculated as the bubble point ofthe overhead

product, light naphtha, under the condenser pressure. Only the temperature calculation in

step g above is required for condenser temperature calculation.

3.2.8 Side-Draw product temperature calculation

In order to calculate the enthalpy ofthe side-draw product which is used in the

calculation ofthe external heat left in the tower, the temperature ofthe stripped side-draw

product is required. The temperatwe ofthe side-draw product can be calculated through

applying an energy balance arovmd the side stripper. It is assumed that each side stripper

has four trays (Watkins, 1979). The vmstripped product enters the first tray from the top

and the steam enters the fourth tray. The stripped side-draw product leaves the side

stripper at the bottom. The stripout and steam exit from the top ofthe side-stripper and

flow back to the atmospheric tower. The calculation follows the steps below:

a. Estimate the steam-free At across the side-stripper using the graph in Watkins (1979).

Polynomial expressions are used to represent the curves in the graph.

b. Estimate the temperature correction for the steam-free At using the correlation in

Watkins (1979). Calculate real At across the side-stripper. This value becomes the

initial guess for At.

c. Assume equal temperature differences across each ofthe four stripping trays

(Watkins, 1979). Calculate the temperature ofthe top tray and the bottom tray

according to the side-draw tray temperature.

d. Make an energy balance around the side-stripper using the temperature calculated

above. The enthalpy of exiting water comes from Moran and Shapiro (1996). The

enthalpy ofthe stripped side-draw product is calculated using this enthalpy balance.

The temperature ofthe stripped side-draw product is obtained from the enthalpy

value.

48

Page 64: 5-95 gap chart

Initial Guess ofthe Temperature ofthe

Side-draw Tray

Start

Enthalpy Calculation of Vapor and Liquid Leaving

the Side-draw Tray

Reflux Heat Calculation Reflux Rate Calculation

Calculation ofthe Molar Fraction ofthe Reflux

Hydrocarbon Partial Pressure CaVs^ation

Side-draw Product Bubble Point Calculation

No

End

Figure 3.11 Flowchart ofthe Side-Draw Tray Temperature Calculation.

49

Page 65: 5-95 gap chart

e. Compare this temperature with the bottom temperature calculated before. If they do

not match, return to step c with the new calculated bottom temperature.

The above iterations are inside the gray block no.4 in the schematic ofthe

atmospheric tower model shown in Figure 3.3.

After the temperature ofthe side-draw product is calculated, the external heat left

in the tower is then calculated by subtracting the heat output ofthe side-draw product and

adding the heat input of steam to previously calculated external heat. This value is then

used in temperature calculation of upper draw trays. In Calculating the temperature

profile in the atmospheric tower, it is assumed that temperature distribution between

adjacent draw trays is linear. Temperature profile along the trays can then be obtained on

the basis ofthe temperature and pressure of draw trays.

3.3 Vacuum Tower Modeling

Under normal pressure and temperature in the flash zone of an atmospheric tower,

the maximum amount of oils which will vaporize is described by the whole crude TBP

cut points between 700°F and 800°F. Further increasing the temperature in the flash zone

will result in thermocracking and coking in the flash zone. The reduced crude usually still

contains a large volume of distillable oil which can be recovered by vacuum process. The

distillable oil is lighter, cleaner, thus, has higher market value than the reduced crude.

The fimction ofthe vacuum tower is to recover this part of distillable oil.

The vacuum tower modeled here is a fuel-oriented type. That is, the vacuum

tower produces feed to FCC unit for gasoline production and produces vacuum residue

for the No.6 fuel oil. Since No. 6 fuel oil has a low market price, the amount of vacuum

residuum should be minimized. This is called pitch operation. If the vacuum residue is

used for asphalt, some gas oil must be left in the residuum in order to provide the proper

degree of plasticity (Watkins, 1979). Pitch operation is used in the refinery considered in

this work. Hence, the same operation mode is considered for the vacuum tower modeling.

A fliel-oriented vacuum tower normally does not require any particular degree of

fractionation between cuts. The HVGO and LVGO side-draw streams produced in a

50

Page 66: 5-95 gap chart

vacuum tower will be combined together after the vacuum tower and become the feed to

FCC unit. Using two side-draw streams is to decrease the liquid and vapor traffic in the

upper section ofthe vacuum tower, which is a common constraint of vacuum tower

operation. A fiael-oriented vacuum tower usually uses pumpback reflux, grid type

contacting sections and chimney draw trays (Watkins, 1979). The pumpback withdraws

liquid on the product side-draw tray, cools the liquid, and return the liquid to the top of

corresponding grid type contacting section. It can be seen from the schematic ofthe

vacuum tower shown in Figure 3.2 that pumpback is used for both separation sections.

There is also an overflash stream in the vacuum tower to provide enough wetting liquid

for the section between overflash liquid draw tray and the flash zone. The overflash can

also be used to control the quality ofthe HVGO, preventing high viscosity material from

entering HVGO. The overflash is set at 2 vol.% ofthe reduced crude according to normal

industrial practice (Watkins, 1979). The overflash stream withdrawn from the vacuum

tower on the overflash draw tray returns to the vacuum crude as the feed to the vacuum

tower.

The light crude processed in the refinery considered in this work does not have a

large volume of heavy ends. Hence, the capacity ofthe vacuum tower is not an active

constraint in operation. In fact, the capacity of downstream units limits the product rates

from the vacuum tower. Therefore, for the range ofthe TBP end point ofthe HVGO,

enough separation capacity is assumed for the vacuum tower.

The modeling of a ftiel-oriented vacuum tower follows the procedure in Watkins

(1979). The basic approach is similar to that used in atmospheric tower modeling.

Therefore, only the steps which are different from the atmospheric tower model are

described here.

3.3.1 Vacuum Tower Furnace Calculation

The outlet temperature ofthe furnace ofthe vacuum tower is set at 763 °F

according to the design documents for the refinery. The reduced crude exiting the

atmospheric tower bottom first exchanges heat with the fresh crude feed and then flows

to the furnace. The inlet temperature is set at 558°F according to the process flow

51

Page 67: 5-95 gap chart

diagram (PFD) ofthe vacuum operation. The heat duty ofthe furnace is calculated as the

enthalpy difference between the reduced crude entering and exiting the furnace.

3.3.2 Side-Draw Product Calculation

The total volume of side-draw products is calculated using the TBP-volumetric

curve ofthe whole crude using the given TBP end point ofthe HVGO. According to

industrial data, the volumetric percent ofthe LVGO in the total VGO is 61.22 vol.% in

the winter and 49.78 vol.%o in the summer. Hence, the volumes of LVGO and HVGO can

be calculated using these values. The volume ofthe overflash is also calculated according

to 2 vol.%) ofthe reduced crude.

3.3.3 Flash Zone Calculation

The flash zone temperature is assumed to be the same as the fiimace outlet

temperature, 780 °F. The flash zone hydrocarbon pressure is calculated based on the

equilibrium flash vaporization. An initial value ofthe flash zone hydrocarbon pressure is

assumed and an equilibrium flash vaporization is performed. The calculated total

vaporization is compared with the total volume ofthe side-draw products plus the

overflash. If two values are different, new value is set for the flash zone pressure. The

secant method (Riggs, 1994) is used in the flash zone pressure solving.

The equilibrium flash vaporization (EFV) curve for the flash zone pressure is

needed to calculate the volume percent ofthe reduced crude flashed. The EFV curve for

atmospheric pressure is obtained from the TBP curve ofthe reduced crude using the

method given by Maxwell (1950). The steps ofthe method are as follows:

a. Construct a distillation reference line using the formulas given below:

c/ fr^i?r TBP{70)-TBP{\0) Slope of DRL = ^f^Z^ ' (^-23)

DRL{50) = TBP{\ 0) + Slope of DRL -(50-10), (3.24)

where

52

Page 68: 5-95 gap chart

DRL- the distillation reference line,

DRL(50)- the 50% ofthe distillation reference line.

b. Construct a flash reference line. The slope and 50%) point ofthe flash reference line

are obtained from those ofthe distillation reference line based on the graph in

Maxwell (1950). The curves in Maxwell (1950) were regressed using polynomial

expression.

c. Use intermediate points to represent TBP curve ofthe reduced crude and

corresponding distillation reference curve. The ATdri, the temperature difference

between the corresponding intermediate points on TBP curve and distillation

reference curve, is calculated using the following equation:

AT^^, = TBP(i) - DRL(i). (3.25)

d. Obtain the ratios between ATDRL and ATFRL, the temperature difference between the

corresponding intermediate points on the EFV curve and the flash reference curve

from the curves given Maxwell (1950). The curves in Maxwell (1950) were regressed

using polynomial expression. ATFRL is then calculated using the following equation:

^Tp^j = AT^^j • Ratio^^. (3.26)

e. Use intermediate points to represent the EFV curve ofthe reduced crude. These

intermediate points are calculated using the following equation:

EFV(i) = FRL(i) + ATp^,. (3.27)

The difference between EFV curves under atmospheric pressure and flash zone

pressure is calculated using the method given by Lin (1988). The steps ofthe method are

given below:

53

Page 69: 5-95 gap chart

a. Find the EFV(50) under the low pressure from the EFV(50) of atmospheric pressure

based on the curves given by Lin (1988). The curves in Lin (1988) were regressed

using polynomial expressions.

b. Use intermediate points to represent the EFV curve under atmospheric pressure.

Usually 11, 9, 7, and 5 intermediate points are used to represent the EFV curve

depending on the requirement ofthe empirical correlations. Calculate temperature

difference among these intermediate points.

c. Assume that the temperature difference among these intermediate points do not

change with pressure. Construct the EFV curve by calculating the intermediate points

ofthe EFV curve under the flash zone pressure using the EFV(50) and temperature

difference obtained in step a and b.

It is claimed that the error of above method is no greater thein 14°C (Lin, 1988).

A constant air leak is assumed for the vacuum tower. The weight of air leak is set

at 44 pounds per hour according to Watkins (1979). The air is considered as an inert in

the calculation. The total pressure ofthe flash zone is then calculated assuming ideal

mixture. The temperature ofthe vacuum residue is assumed to be the flash zone

temperature.

An energy balance is carried out around the flash zone. The enthalpy of air is

obtained from Moran and Shapiro (1996). The external heat left in the tower, which is

enthalpy ofthe vapor going upward, is calculated. The flash zone calculation of vacuum

tower is similar to that of atmospheric tower.

3.4 Separation Section Calculation

The vacuum tower has three separation sections: Overflash section, HVGO

section and LVGO section. There is no overhead condenser. The cooling is provided

totally by the pumpbacks.

It is assumed that 3.81 millimeter Hg pressure drop for each chimney tray and

each grid separation section according to plant data. The pressure distribution is then

calculated from the assumed pressure drop and the flash zone pressure calculated above.

54

Page 70: 5-95 gap chart

The energy and material balances calculation ofthe separation sections in the vacuum

tower are similar to those in the atmospheric tower model.

3.5 Comparison between the Developed Model and Tray-to-Tray ChemCad Model

Since the crude unit model is a simplified model, it is necessary to compare it

with a more rigorous model to find out if the developed model can satisfy the

requirement ofthe refinery-wide optimization. For this purpose, a ChemCad model was

constructed for comparison. The ChemCad model is a rigorous tray-to-tray model

constructed using software package CHEMCAD 4.0. In the ChemCad model, 45 pseudo-

components are used to represent the crude feed. The atmospheric tower and vacuum

tower in ChemCad model have exactly the same design as the developed model. Same

crude is used in the ChemCad model in the Summer Mode and the Winter Mode. The

values ofthe overhead pressure, pressure drop and steam usage used in the developed

model are also used in the ChemCad model. The side-draw tray locations and the

sidestripper type are also the same. In addition, the heat duties ofthe pumparounds and

condenser for both models are the same. The fiowsheet of ChemCad model is shown in

Figure 3.12.

The most important fimction ofthe crude unit is predicting the volumetric flow

rates ofthe side-draw products. The volumefric flow rates ofthe side-draw products are

decided by the given ASTM 95%) point or TBP 100%. The separation specifications of

side-draw products are shown in Table 3.7. The same specifications are used in the

ChemCad model except that the ASTM 90%) point is used for Heavy Vacuum Gas Oil.

The discrepancy may be due to the fact that the Crude Assays obtained from the refinery

do not have detailed information at the high boiling range which results in the inaccurate

information ofthe pseudo-components at the heavy ends ofthe distillation curve.

55

Page 71: 5-95 gap chart

0) • 0

o 2

C D

0) •0 3 H U

•a u e tt)

u 14-1

o 0

•H 4-> IB

e o

0)

u

•H [I4

Page 72: 5-95 gap chart

Table 3.7 Separation Specifications of Side-Draw Products.

Side-draw Product

Light Naphtha

Heavy Naphtha

Light Distillate (Kerosene)

Heavy Distillate (Diesel)

Heavy Vacuum Gas Oil

Specification, °F

257

430

530

650

1050

SpecificationType

ASTM 95%

ASTM 95%

ASTM 95%

ASTM 95%

TBP 100%

The comparisons between the side-draw product flow rates ofthe developed

model and ChemCad model in the Summer Mode and the Winter Mode are shown in

Figure 3.13 and Figure 3.14. The flow rates match well for most ofthe side-draw

products. The small mismatch will only affect the refinery-wide optimization if the

separation specifications, ASTM 95%) or TBP 100%) point, are at the specification limits,

which is rarely the case in the optimization studies. The side-draw product rates are

usually constrained by the capacities of downstream units instead ofthe specification

limits.

Another important test is the gain comparison, that is, the comparison between the

changes of product flow rate versus the changes ofthe separation specifications obtained

from two models. Having correct gain is critical in optimization studies. The optimization

routine needs the gain information between the decision variables and the optimization

objective fianction to decide the optimization search direction. The gains between the

separation specifications and the product rates significantly affect the gain between the

separation specifications and the objective function. Hence, the developed model must

have correct gains so that it can be used in the following optimization studies.

The normal operating values for the base case are listed in Table 3.7. The furnace

ouflet temperature is set at 653°F in the base case. Each ofthe ASTM 95% point, TBP

100% point, and furnace outiet temperature is perturbed by plus 5°F and minus 5°F

sequentially in the test.

57

Page 73: 5-95 gap chart

16000

14000

12000

^ 10000 (3

2 8000 oa

® 6000

4000

2000

0

D Simplified Model

• ChemCad Model

m light heavy kerosene diesel

naphtha naphtha AGO VGO vacuum

residue

Figure 3.13 Comparison of Simplified model and ChemCad Rigorous Model, Summer Mode.

18000

16000

14000

12000

I" 10000

m 8000 CQ

6000

4000

2000

0

D Simplified Model

• ChemCad Model

M light heavy kerosene diesel AGO VGO vacuum

naphtha naphtha , „, residue volume%)

Figure 3.14 Comparison of Simplified Model and ChemCad Rigorous Model, Winter Mode.

58

Page 74: 5-95 gap chart

The product rates after the perturbation were compared to those ofthe base case

and the gains were estimated by the formula given below.

^ ^ - ^ > . .

^^y Ti-Tj,^,, (3.28)

The average gains were calculated as the arithmetic average of gains ofthe

positive and negative perturbation.

The comparison between the Gains obtained from the developed model and the

ChemCad model are shown in Table 3.8. It is observed that only diagonal elements have

significant values. It means that the ASTM 95% or TBP 100% for one product mainly

influences the volumetric flow rates of this product and adjacent heavier product while

having little effect on the volumetric flow rates of other products. It is also observed that

although there is some mismatch between the gains ofthe developed model and the

ChemCad model, the corresponding gains are in the same order and have the same sign.

The mismatch is expected because the developed model uses simplified approaches and

empirical correlations. However, the mismatch is not expected to distort the searching

direction in optimization.

Since the volumetric fiow rates ofthe side-draw product are the most important

in the refinery wide optimization, it is also necessary to find if other operating variables

have significant influence on the product rates. Another process variable which may have

significant influence is the pumparound heat duty. The pumparound heat duties are

defined explicitly in the atmospheric tower modeling. In the vacuum tower, the

pumparound duties are defined by the returning temperature.

59

Page 75: 5-95 gap chart

o CL,

^ o^ i n ON

s H OJ < . ^ u

-a o B

X ! rt o B u

.£3

u J5 • * — •

• ^

del

o

T3 U

s c u

.13 +-•

g o

^4—(

-o

'3 -> .£i

u x) o s ;-! u s H P

00

_o E o

-T3 O

OH IZl

3 on i-i U > +-; _g "o

T3 C2

o m CO

_c '3

J 3

t 4 - l

O

o w

'S Cu B o U

00 rn u

i

OH

m O o > -a §

r\

0)

s t i

lU

! • (L» H -t-» (U

"3 H O

c

<1

I f^"" ^

o o o

9 o o

d o ^

C50

i n

fa fa

^

T3 O

T 3

U

a U

^

O

U

a >< o O H

0 0

d f^ I o ^

>o oo J2 ^ ^ s i n

Ci3

9 9 9 i-J H J hJ PQ 03 CQ m m OQ

o B

u S u

O

oo

VO

00 ON

9 OQ DQ

(L)

o

ON ON

m

o

d 00

OS

fa fa

9 OQ OQ

'O 1) 13 S o

<u T3 O B

-o u a o

.a .s -§ fa fa fa < < <

fa fa < <

vo ON

vo

00 ON

H 2 vo 1

0 0

VO

i n

d <N

VO ON

Q

OQ CQ

p^

a

OQ OQ

-a o

-a u

a >< o

2 S

O

O

a U

t ^ i n

o 00 (N VO

fa <

fa

OQ OQ

-a o a u 15

o i-i

-a' fa <

oo d

Q

OQ CQ

O

a u a

.£5

u o" GO c3

fa <

i n 00 i n

i n

fa fa

^

OQ CQ

o

<U

Id

a o ;-! O H o, < 0 0 CO

fa <

o o r<-i

i n

i n vo

(N 00

m vo vo ON

o

ON

fa o

Q

OQ CQ

o

a T 3 Ci3 U a u 00 >

fa <1

i n od i n

fa o I' OQ CQ

-a o

• T 3 (U H-»

CS

a >< o kH

O H

< d 00 > fa <

o p

i n vo

ON

o ON

o d

00

o

ca Q

OQ OQ

-a o a u a u

CO

fa < ]

r<-i

i n

fa o

Q

CQ CQ

u -a o a -a (U H-t CO

o O , O H

<

CO

LH

fa <

60

Page 76: 5-95 gap chart

The results ofthe corresponding sensitivity tests on pumparounds are shown in Table 3.9

and Table 3.10. It is found that the pumparound heat duties have little effect on the

product flow rates as long as the pumparounds satisfy the cooling requirements of each

tower and provide enough reflux in each separation section. Pumparounds may affect the

energy usage in the crude unit. Nevertheless, energy consumption is a secondary factor in

the refinery-wide optimization. Hence, the pumparound heat duties are fixed when

refinery-wide optimization is conducted and they are not used as decision variables. The

same approach of selecting decision variables was used in other units introduced in the

following text. Those process variables only affecting energy consumption are not

selected to be decision variables ofthe refinery-wide optimization.

Table 3.9 Comparison between the Gains Obtained from the Simplified Model and ChemCad Model, Pumparound Heat Duty versus Product Flow Rate, Atmospheric Tower, Summer Mode.

AFin, ChemCad model

AFjn, Approximated model

AFhn > ChemCad model

AFhn, Approximated model

AFid. ChemCad model

AFjd, Approximated model

AFhd: ChemCad model

-^Fhd' Approximated model

AFago, ChemCad model

AFago, Approximated model

Units

BBL/hr/MMBtu

BBL/hr/MMBtti

BBL/hr/MMBtu

BBL/hr/MMBtu

BBL/hr/MMBtu

BBL/hr/MMBtti

BBL/hr/MMBtu

BBL/hr/MMBtti

BBL/hr/MMBtu

BBL/hr/MMBtti

AF

l^Qpa^

-0.195

0

0.3749

0

-0.0993

0

0.1097

0

-0.0035

0

AF

^Qpal

0.0335

0

-0.4677

0

-0.4229

0

0.8632

0

-0.0008

0

61

Page 77: 5-95 gap chart

Table 3.10 Comparison between the Gains Obtained from the Simplified Model and ChemCad Model, Pumparound Rettim Temperature versus Product Flow Rate, Vacuum Tower, Summer Mode.

Units

APV

^Tpal

APV

^Tpa2

AFhvgo, ChemCad model

AFhvgo, Approximated model

AFresid, ChemCad model

AFresid, Approximated model

BBL/hr/°F -0.0009

BBL/hr/°F 0

BBL/hr/T 0.0008

BBL/hr/°F 0

-0.1765

0

0.0327

0

On the other hand, pumparounds may become the decision variables of single-unit

optimization ofthe crude unit. This is because energy usage is a major factor in single-

unit optimization. The pumparounds are used to preheat the crude. While satisfying

separation requirement and heat requirement in the atmospheric tower and vacuum tower,

maximizing pumparounds may be beneficial since it provides maximal heat to preheat the

crude, which may decrease the fuel usage in the preheat furnace.

Another comparison is done on the temperature profile ofthe atmospheric tower

ofthe developed model and ofthe ChemCad Model. The comparison is on the basis of

using the same separation specifications, condenser pressure, pressure drop in the tower

and the fiimace outlet temperature. The comparison ofthe temperature distribution in the

crude tower between the developed model and the Chemcad model is shown in Figure

3.15.

62

Page 78: 5-95 gap chart

800

700

600

500

2 400 PQ OD

300

200

100

0

— Simplified Model

- ChemCad Rigorous Model

10 15 20 25 tray number

30 35 40

Figure 3.15 Comparison of Temperature Profile in the Atmospheric Tower.

It is observed that although the crude unit model in this work is a simplified

model and linear temperature distribution is assumed in each separation section,

temperature profiles in two models are similar in shape. The biggest temperature

difference appears on tray 29 to 35, where the temperatures predicted by the simplified

model are about 20%) higher than those of ChemCad model. There are only small

differences from tray 1 to tray 28, where the average relative difference is 2.9%. The

average relative temperature difference for all trays is 5.8%. Tray temperature difference

only affect the estimation of utility cost in the atmospheric tower, which only has an

insignificant effect on plant-wide economics.

3.6 Auxiliary Processes in the Crude Unit

There are some auxiliary processes in the crude unit which prepare the side-draw

products exiting the atmospheric tower and vacuum tower for ftirther processing in the

refinery. These processes includes Rerun unit. Residue Oil Solvent Extraction (ROSE)

unit, debutanizer, Naphtha Splitter, etc. These units are not critical from plant-wide

63

Page 79: 5-95 gap chart

operation point of view. However, they must be included in the refinery-wide model in

order to run the refinery-wide model. The models of these units are required when

calculating the amount ofthe feeds to downstream units or products for sale. In addition,

the capacities of these processes may become active constraints in refinery-wide

optimization.

Simplified models were developed for these processes. Since the operations of

these processes in the Summer Mode are quite different from the operations in the Winter

Mode, different parameters were used to account for the difference in operation.

3.6.1 Rerun Unit

The fianction ofthe Rerun Unit is to recover the diesel components from the AGO

and the LVGO. Due to incomplete separation between the diesel, the AGO and the

LVGO, the AGO stream and the LVGO stream may contain some components whose

boiling points are in the diesel boiling range. Since diesel itself has high market value, it

is more profitable to recover this part of diesel component and sell it directly instead of

using it as the feed to FCC unit to make FCC gasoline.

The core ofthe rerun unit is a distillation column. Using the AGO stream and part

ofthe LVGO stream as feed, the diesel components come out from the column overhead.

The overhead stream combines with the diesel directly produced from the atmospheric

tower and flows to hydrodesulftirizer (HDS) unit. The bottom stream is combined with

the rest ofthe LVGO stream and the HVGO stream as the feed to the FCC unit.

All AGO side-draw product from the atmospheric tower is sent to Rerun unit.

Only part ofthe LVGO stream from the vacuum tower is sent to the rerun unit. The

typical percent of LVGO used as Rerun feed is 13 vol.%) in the Summer Mode and 11

vol.% in the Winter Mode. The volumetric flow rate ofthe overhead product ofthe rerun

distillation column is about 25 vol.%) ofthe total rerun feed in the summer and 23 vol.%)

ofthe total rerun feed in the winter. These values are used directly in the model to

calculate the amount ofthe LVGO stream sent to the rerun unit and the amount ofthe

diesel component produced in the Rerun unit.

64

Page 80: 5-95 gap chart

3.6.2 Residue Oil Solvent Extraction (ROSE) Unit

The function ofthe ROSE unit is to recover FCC cracking feed from the vacuum

residue. The vacuum residue is rich in metals, tar, and asphalt. Metals are poisonous to

the FCC catalyst. Tar and asphalt can deposit large amount of carbon on the FCC catalyst

which could severely decrease the catalyst activity and could lower the FCC unit

processing capability. Hence, the vacuum residue can not be used directly as FCC feed.

Without treatment, vacuum residue can only be sold as low-value No. 6 fuel oil.

However, the vacuum residue still contains a large amount of hydrocarbons that is

suitable for catalytic cracking. If this portion of hydrocarbons in the vacuum residue can

be recovered, it can be used to produce high-value FCC gasoline. Thus, recovering FCC

feed from residue is economically favorable.

Since the vacuum residue has very high boiling range, distillation can not be used

to separate the vacuum residue. The most efficient way to separate the vacuum residue is

solvent extraction. The solvents normally used in a refinery are propane and butane.

Paraffinic and naphthenic hydrocarbons have a large solubility in the above mentioned

solvents while metals and asphalt material have very low solubility. The extraction

operation consists of a series of countercurrent contacts between oil and solvent (Nelson,

1958). The separation between the solvent and the hydrocarbon is easily done by flash

because the solvent is much lighter than the hydrocarbon. The recovered hydrocarbon is

called deasphalted oil (DAO).

According to the refinery considered in this work, the entire vacuum residue is

sent to the ROSE unit. The typical percent of DAO product is about 60 vol.%) ofthe

vacuum residue in both the Summer Mode and the Winter Mode. This value is directly

used to calculate the DAO rate from the vacuum residue. DAO is then considered as a

crude cut and its properties are calculated as the same as those of any other crude cut.

DAO is combined with the HVGO produced directly from the vacuum tower and rerun

column bottom stream as the feed to FCC unit. The insolvable part of vacuum residue is

sold as No. 6 fuel oil.

65

Page 81: 5-95 gap chart

3.6.3 Debutanizer

The function ofthe debutanizer in the crude unit is to separate the light

components from the light naphtha in order to stabilize the light straight-run gasoline

(LSR). Most ofthe overhead products from the debutanizer are propane and butane. This

stream becomes liquefied petroleum gas (LPG) and flows to the gas plant ofthe refinery.

The bottom product from the debutanizer flows to the naphtha splitter to make LSR and

reformer feed. The Debutanizer adds flexibility to refinery operation. When the Reid

Vapor Pressure (RVP) specification for gasoline is high in the winter, more butane is left

in the LSR. When the RVP specification is lower in the summer, most ofthe butane

enters LPG and is used as alkylation feed to produce alkylate, a lower RVP gasoline

blending stock.

In the atmospheric tower model, the vapor stream and liquid stream coming out

the overhead condenser are combined as the light naphtha stream. Before the calculation

ofthe debutanizer, the amount of light gas in the light naphtha is calculated according to

industrial data and is subtracted from the light naphtha

The LPG production rate from the debutanizer depends on the quality ofthe

crude. Since the crude slates are different in different seasons, the LPG product rates are

different. Using industrial data, the weight and volume of LPG are calculated from the

weight and volume ofthe debutanizer feed. The weight ofthe debutanizer bottom steam

is then estimated according to material balance. According to industrial data, the total

volume loss is negligible in the debutanizer, less than \%. In other words, the summation

ofthe volumetric flow rates of LPG and debutanizer bottom product is almost equal to

the volumetric flow rate ofthe debutanizer feed. Assuming that there is no volume loss,

the volumetric flow rate ofthe debutanizer bottom stream is then calculated.

3.6.4 Naphtha Splitter

The function ofthe naphtha splitter is to split the debutanizer bottom stream into

light straight-run (LSR) gasoline and heavy naphtha. The naphtha splitter also adds

flexibility in refinery operation. The operation ofthe naphtha splitter can be adjusted

66

Page 82: 5-95 gap chart

according to downstream gasoline blending. If the octane number ofthe gasoline pool is

low, more light naphtha is used as the reformer feed to produce high-octane reformate.

When RVP specification of gasoline is low in the summer, the amount of high RVP light

straight-run gasoline (LSR) is reduced. More naphtha is used as reformer feed to produce

medium-RVP reformate.

Using industrial data, the volumetric flow rates ofthe LSR are calculated from the

volumetric flow rate ofthe feed to naphtha splitter. The LSR is then considered as a

crude cut and its properties, including weight, are calculated as any other crude cut. The

weight ofthe splitter bottom is estimated based on material balance. According to

industrial data, the total volume loss is negligible in the naphtha splitter and is less than

1%). Hence, assuming that there is no volume loss, the volumetric flow rate ofthe

debutanizer bottom stream is calculated from the volumefric flow rates ofthe feed and

the LSR.

67

Page 83: 5-95 gap chart

CHAPTER 4

FCC FEED CHARACTERIZATION AND

MODEL BENCHMARKING

4.1 Process Overview

Fluidized Catalytic Cracking, better known as FCC, is one ofthe most important

and complicated processes in petroleum refining. For many refiners, FCC is the key to

profitability in that the successful operation ofthe unit can determine whether or not a

refiner can stay in business and remain competitive in today's market (Sadeghbeigi,

1995). In 1989, 40 vol.% of gasoline blending stocks in the U.S. were produced from the

FCCU (Lin, 1993). In the refinery considered in this work, FCC provides about 45 vol.%

blending stock ofthe gasoline pool. Since gasoline is the largest volume product from a

fuel-oriented refinery, an economic study on the refinery-wide operation must consider

the operation ofthe FCC unit.

The FCC unit converts large quantities of heavy feed into more valuable product,

called FCC gasoline. The FCC unit also produces light cycle oil (LCO), heavy cycle oil

(HCO) or slurry oil, and light gas. LCO has the same boiling range as of diesel. It is

blended with the diesel stream directly from the crude unit to produce the diesel product.

HCO from FCC is either recycled to the FCC feed or combined with the vacuum residue

from the crude unit to make No. 6 fuel oil. In the refinery considered in the study, most of

the HCO is used to make No. 6 fuel oil. Hence, the HCO is not recycled to the FCC feed

in the model. Light gas produced in the FCC unit is the main resource of light gas in the

refinery. Light gas is sent to the gas plant to produce fuel gas, liquidified petroleum gas

(LPG), propylene, and butane. Light gas also provides the feedstock to the alkylation unit

in the refinery.

From the block flow diagram ofthe refinery in Figure 3.1, it can been seen that

the FCC unit receives atmospheric gas oil (AGO), light vacuum gas oil (LVGO), heavy

vacuum gas oil (HVGO) and deasphalted oil (DAO) from the crude unit. In many

refineries, FCC feed goes through a hydrotreater before it enters the FCC unit if the FCC

68

Page 84: 5-95 gap chart

feed has a high content of sulfur and metals. Sulfur in the feed may enter the flue gas of

the FCC unit to cause environmental pollution. Sulfiir also enters the product streams of

the FCC unit such as gasoline, LCO, and HCO that have upper limits on sulfur content in

the products. Metals in the FCC feed are catalyst poisons. Since the crude oils processed

in the refinery considered in this work belong to light, low sulfur crude type, the FCC

feed does not have high content of sulfur and metals. Therefore, a hydrotreater for the

FCC feed is not required in this particular refinery.

The FCC unit in the refinery considered in the work is a Model IV FCC unit. The

flow diagram of a Model IV FCC unit is shown in Figure 4.1. The gas oil from the crude

unit is first preheated in a fiimace and then mixed with the regenerated catalyst and enters

the reactor riser. Most ofthe cracking reaction occurs in the riser. The residence time of

the gas oil in the riser is about 2 seconds (Sadeghbeigi, 1995). Short contacting time is

required to prevent secondary cracking that cracks high-value products such as gasoline

into light gas. The catalyst is separated from the product vapor in the disengaging zone on

the top ofthe reactor and flows to the regenerator. The coke deposited on the catalyst is

bumed in the regenerator. After regeneration, catalyst regains its activity and flows to the

riser again. The product vapor exits the reactor from the top and enters the main

fractionator. In the main fractionator, the mixed product stream is separated into several

products. The liquid stream withdrawn from the overhead condenser is FCC gasoline.

The vapor steam from the overhead condenser flows to the wet gas compressor and on to

the gas plant for ftirther treatment. LCO is withdrawn from a side-draw tray ofthe tower.

After steam stripping, LCO flows to the diesel storage tank. HCO or slurry is withdrawn

from the bottom ofthe main fractionator and the bottom ofthe sidestripper. Most of HCO

or slurry is used as the feedstock for No. 6 fuel oil.

69

Page 85: 5-95 gap chart

G o 00

c o c« ; -< u o

*H—»

13 13 O T3 <U N

•-3

fa >

O

O o

a o

C/3

3 GO

70

Page 86: 5-95 gap chart

The distinguishing feature of a model IV FCCU is that it does not have a slide

valve on the catalyst circulation line like most of modem FCCU does. Hence, there is no

direct control of catalyst circulation rate between the reactor and regenerator. Only

indirect control is possible by varying the driving force for catalyst flow (Huq and

Morari, 1995). The model IV FCCU also has low metallurgical limit, about 1000 °F for

the riser temperature, while modem FCCUs may have the riser temperatures as high as

1050 °F(Arbeletal., 1995).

4.2 Model IV FCC Unit Modeling

McFarlane et al. (1993) developed a dynamic Model IV FCC unit model. The

model uses a continuous stirred-tank reactor (CSTR) for the riser and does not include a

yield model. Ellis(1996) developed a steady-state model based on McFarlane's dynamic

simulator. Plug-Flow Reactor (PFR) model is used to describe the flow and reaction in

the reactor. The ten-lump yield model proposed by Jacob et al.(I976) is used in the

steady-state model to calculate the product distribution. This model characterizes the gas

oil feed on the basis ofthe molecular stmcture and the molecular weights. The ten-lump

reaction network is shown in Figure 4.2. The boiling ranges ofthe individual lumps are

shown in Table 4.1.

Among the ten lumps, the lump of aromatic substituent molecules represents the

branches ofthe aromatic rings. Since they have completely different cracking

characteristics from the aromatic rings that they attach to, they form separate lumps. The

aromatic substituent group can be paraffin chains or naphthenic groups (Jacob et al.,

1976).

The FCC feed components with boiling points (I) greater than 650°F and (2)

between 430 °F and 650 °F are each grouped in four lumps: paraffins, naphthenics,

aromatics, and aromatic substituent groups. All eight lumps are cracked into gasoline,

and light gas in the riser. Four lumps with boiling point greater than 650°F compose

HCO. Four lumps with boiling point between 430°F and 650°F compose LCO. Gasoline

contains Cs's up to a boiling point of 430°F. The "C" lump contains coke and light gases

71

Page 87: 5-95 gap chart

(C1-C4). Gasoline lump and "C" lump are the products ofthe cracking reactions. The

kinetic models used for this reaction network were taken from Jacob et al. (1976) and

Arbel et al. (1995) and assume first-order reactions with Arrhenius rate constants

for each ofthe 20 reactions between the ten lumps as shown in Figure 4.2.

I - PI

Ph - "

NI

Nh -I

As! -1

I-Ash

Arl —,

Asl

Figure 4.2 Ten-Lump Reaction Network.

72

Page 88: 5-95 gap chart

Table 4.1 Boiling Range of Lumps in Ten-Lump FCC Reaction Network.

Lump Description Boiling Range

Ph Wt. % paraffinic molecules

Nh Wt. % napththenic molecules

Ash Wt. % aromatic substituent molecules

Arh Wt. % carbon atoms among aromatic rings

Pi Wt. % paraffinic molecules

Ni Wt. % napththenic molecules

Asi Wt. % aromatic substituent molecules

Ari Wt. % carbon atoms among aromatic rings

G Wt. % gasoline

C Wt. % coke and

1. Butane 4. Propene

2. Isobutane 5. Propane

3. Butane 6. Gases<C2

T>650°F

T>650°F

T>650°F

T>650°F

430°F<T<650°F

430°F<T<650°F

430°F<T<650°F

430°F<T<650°F

C5's-430°F

(C1-C4, and coke)

4.3 FCC Feed Characterization

Feed information is an input to the FCC model. Feed information required by the

FCC model includes volumetric flow rate, API gravity, and the weight fractions ofthe

eight lumps in the feed.

4.3.1 Volume and Weight of FCC Feed

FCC feed is composed of AGO from the rerun unit bottom, LVGO, HVGO, and

DAO. The volume and weight of these four streams have been calculated in the cmde

unit model described in Chapter 3. The volume and weight ofthe FCC feed is the

summation of these of four streams:

^FCC ~ ^.-iGO "'" ^LVGO "*• ''^HVGO + ''^DAO ' (4.1)

73

Page 89: 5-95 gap chart

^fCC = ^AGO + WUGO + KvGO + ^OAO > (4-2)

where

Vpcc, VAGO, VLVGO, VHVGO, VDAQ- volumetric flow rates of FCC feed, AGO, LVGO,

HVGO, and DAO, respectively, BBL/D,

WFCC. WAGO, WLVGO, WHVGO, WDAO- weight flow rates of FCC feed, AGO, LVGO,

HVGO, and DAO, respectively, MLB/D.

The API gravity ofthe FCC feed is calculated using the weight and the volume of

the FCC feed using the formula below:

141 5 APIpcc= • 13^-5, (4.3)

(Wp,.,. • 0.454)/(Vp^c-0.159)

where

APIFCC- API gravity ofthe FCC feed,

WFCC- weight flow rate ofthe FCC feed, MLB/D,

Vpcc- volumetric flow rate ofthe FCC feed, BBL/D.

4.3.2 FCC feed Characterization

To use the ten-lump yield model, the weight percentages ofthe lumps in the FCC

feed must be known. Only the first eight lumps are in the feed: Ph, Nh, Ash, Arh, Pi,N|„

Asi,and Ari Lump gasoline and lump "C" are not in the feed.

Some industrial data of FCC feed obtained from the refinery considered in this

work are listed in Table 4.2. Although the given data can not be directly used to calculate

the weight percentage ofthe eight lumps in the FCC feed, at least they provide the

general composition ofthe FCC feed which can be used to validate the correlations used

in the feed characterization.

74

Page 90: 5-95 gap chart

Tabel 4.2 Industrial Data of FCC Feed.

Feed Property Plant Data, Typical FCCU Combined Fresh Feed Analysis, May through August, 1998

Aniline point, °F 197-224

Viscosity, centipoise 6.4

Refractive Index, 60°F 1.5067-1.5082

API 22.0-23.3

Light Components, <650°F 10% TBP, 606-622°F

20% TBP, 678-694°F

Paraffins 56.9-59.1 (vol.%)

Naphthenes 22.2-24.8 (vol.%)

Aromatics 18.1-18.7 (vol.%)

To make the comparison between the feed characterization and the limited

industrial data, the operating conditions in the cmde unit must be the same for the model

and as for the industrial data because the operation of cmde unit significantly affects the

properties ofthe FCC feed. Since the given data for the FCC feed is for the Summer

Mode Operation, typical operating conditions in the summer are used. They are the same

as the normal values ofthe operating variables listed in Table 3.7 in Chapter 3.

The weight percent ofthe four light lumps in the FCC feed can be determined

from the boiling point ofthe light lumps. The light lumps all have a boiling point less

than 650°F. Among the four streams of FCC feed, AGO is the lightest and DAO is the

heaviest. The portion ofthe FCC feed on the whole cmde TBP curve begins with the IBP

of AGO and ends with EBP of DAO. From the whole cmde TBP-volumetric curve, the

cmde volume with the boiling point between the IBP of AGO and 650 °F is obtained.

The volume fraction of light lumps in the total FCC feed is then calculated as below:

VT _ "^lighilump

'^ light lump jrj. ' y.^-^)

y i fcc

75

Page 91: 5-95 gap chart

where

Viight lump- volume fraction ofthe light lumps in the FCC feed,

VTiight lump- volumetric flow rate ofthe four light lumps in the FCC feed, BBL/D,

VTFCC- volumetric flow rate ofthe FCC feed, BBL/D.

The API gravity ofthe light lumps is estimated from the API-volumetric curve ofthe

whole cmde. Then, the specific gravity of light lumps is calculated. The weight fraction

ofthe light lumps in the FCC feed is calculated as blow:

^ _SPG„^,-Vi,i,ii.np-01S9

""""""" WTp,,-0.4536 ' ^ ^

where

SPGiight lump- specific gravity ofthe light lumps in the FCC feed,

Wiightiump- weight fraction ofthe light lumps in the FCC feed,

WTFCC- weight flow rate ofthe FCC feed, MLB/D.

To calculate the weight percentages ofthe eight lumps in the feed, the weight

percentages of paraffins, naphthenes, and aromatics (PNA) in the FCC feed must be

known. The TOTAL method (Dhulesia, 1986) and n-d-M method (ASTM, 1985) are

used here.

TOTAL method is used to calculate the refractive index (RI) at 20°C and 60°C.

TOTAL method is also used to calculate the fraction of carbon in the aromatic ring in the

FCC feed. It is claimed that TOTAL method is more accurate than other methods in

predicting refractive index (Dhulesia, 1986) and aromatic carbon content (Sadeghbeigi,

1995). The correlations in the TOTAL method were developed from multiple linear

regression of 33 feedstocks (Dhulesia, 1986).

RI is an important index of FCC feed. It shows how refractive or aromatic a FCC

feed is. The higher the RI, the less crackable the FCC feed is (Sadeghbeigi, 1995). The

correlations of TOTAL method are given below:

76

Page 92: 5-95 gap chart

RI(20) = 1 + 0.8447-SPG""'' -(VABP^ + 273.16)'"'"'' -MW'"""'', (4.6)

RI(60) = 1 + 0.8156 • SPG"''' -(VABP^ + 273.16)-""''' -MW'""""', (4.7)

where

RI (20)- refractive index at 20°C,

RI (60)- refractive index at 60°C,

SPG- specific gravity of FCC feed,

VABP°c- volumetric average boiling point, °C,

MW- molecular weight of FCC feed.

For consistence reason, the molecular weight ofthe FCC feed used in the

equations 4.6 and 4.7 are also calculated from the correlation in the TOTAL method as

below:

MW = 7.8312 -10-' - SPG'""''' - VABPc'"'' • AP"'"', (4.8)

where

AP- Aniline point, °C.

VABP is calculated from the ASTM boiling curve ofthe FCC feed. The ASTM

boiling curve is converted from TBP curve of FCC feed. Please refer to Chapter 3 for the

details ofthe calculations. The aniline point ofthe FCC feed uses the average value ofthe

aniline point ofthe FCC feed in industrial data, 99.2°C. The moles ofthe FCC feed is the

summation ofthe moles of AGO, LVGO, HVGO, and DAO. The molecular weight of

FCC feed is then calculated using the moles and the weight ofthe FCC feed.

The RI (20) and RI (60) calculated in equations (4.6) and (4.7). Only RI (60) of

the FCC feed is measured in the refinery considered in the work. The RI (60) predicted

by the TOTAL method, 1.49982, is just a little smaller than the normal range of FCC

feed 1.5067-1.5082. However, the n-d-M method which is used to calculate the

77

Page 93: 5-95 gap chart

naphthene content and the paraffin content ofthe FCC feed is very sensitive to the value

of refractive index (RI). Small differences in RI may cause a large error in PNA

prediction when using the n-d-M method. It was reported that a 35%) drop in the aromatic

content was observed when the RI (20) dropped from 1.5105 to 1.5000 (Sadeghbeigi,

1995). Hence, accurate values of RI(20) and RI(60) are required. Here, a bias is used to

adjust the calculated RI(60). The bias between the average value of RI (60) ofthe FCC

feed in the industrial data, 1.50745, and the calculated RI (60) of normal operating

condition is calculated as below:

dRI(60) = 1.50745 - RI(60)„„^,^,, (4.9)

where

RI(60)normai- Calculated RI(60) under normal operating conditions using TOTAL method,

dRI (60)- the bias of RI(60).

The dRI (60) calculated here is then used in adjusting the calculated RI (60)

thereafter as shown below:

RI(60)^,j^,^, = RI(60),o,^ +dRI (60), (4.10)

where

RI(60)adjusted- adjusted value of RI(60),

RI (60)TOTAL- refractive index at 60°C calculated by TOTAL method.

The TOTAL method and the n-d-M method use the value of RI (20) in

composition calculations instead of RI (60). Now the question is how to estimate the

difference between RI (20) and RI (60). It is assumed that TOTAL method is consistent

in predicting RI (60) and RI (20) for one feed. Hence, the RI (60) and RI (20) calculated

from the TOTAL method is used to estimate the difference between RI (60) and RI (20).

The final value of RI (20) is calculated using the following formula:

78

Page 94: 5-95 gap chart

RI(20)^,,..,,,, = RI (60),,j^,,, - (RI(60),orAL - Rl(20)rorAL ) , (4-11)

where

RI(20)adjusted- adjusted value ofthe RI(20),

RI (20)TOTAL- refractive index at 20°C calculated by TOTAL method.

The correlation in TOTAL method is used to calculate the aromatic carbon

content in the FCC feed. The correlation has the formula as below:

%C, = -814.136 + 635.192 • RI(20)^,j^^,^, -129.266 - SPG + 0.013 • MW

-0.340- SUL + 0.757 In(VIS)

where

%)CA- aromatic carbon content ofthe total carbons in the FCC feed, wt.%),

SPG- specific gravity ofthe FCC feed,

SUL- sulftir content ofthe FCC feed, wt.%,

VIS- viscosity of FCC feed, centistoke.

The viscosity in the industrial data has the units of centipoise. The conversion between

centistoke and centipoise uses the following formula:

, centipoise centistokes = . (4.13)

density

The n-d-M method is an ASTM (D-3238-85) method used to estimate the carbon

distribution in aromatic ring stmcture, naphthenic ring stmcture, and paraffin chains.

Only the correlation of refractive ring, which includes aromatic ring and naphthenic ring,

is used here. The formulas are as follows:

V = 2.5-[RI(20)-1.4750\-(SPG,O^^ -0.8510), (4.14)

79

Page 95: 5-95 gap chart

m = (SPG,o,c -0.8510)-1.11-[RI(20)-I.4750] , (4.15)

where

V, Tu- n-d-M method parameters,

SPG20''c- specific gravity at 20°C.

If T!j is negative: %C« =820-m-3-SUL + ^ ^ ^ , (4.16) MW

If tu is positive: %C« = 1440-m-3-SUL + ^^^ , (4.17) MW

where

% C R - weight percentage ofthe carbons of refractive ring stmcture in the total carbons in

the FCC feed, wt.%.

Using the aromatic carbon content calculated in equation 4.12, the naphthenic

carbon content c£in be calculated as below:

%C^=%C^-%C^ (4.18)

%Cp=100-%C^ (4.19)

where

% C N - weight percentage ofthe carbons of naphthenic rings in the total carbons in the

FCC feed, wt.%,

%)Cp- weight percentage ofthe carbons of paraffin chains in the total carbons in the FCC

feed, wt.%.

CA, CN, and C? calculated above are the weight percentage ofthe carbon atoms in

a certain molecular structure in the total weight of carbon atoms. The weight percentage

of aromatic rings, naphthenic rings, and paraffin chains are calculated next. For aromatic

ring, the carbon/hydrogen ratio is 1:1 while the carbon/hydrogen ratio in naphthenic ring

is 1:2. Considering the fact that the paraffin molecules in the FCC feed are big molecules

80

Page 96: 5-95 gap chart

with long paraffin chains, most of carbon atoms are in the middle of paraffin chains.

Hence, carbon/hydrogen ratio in the paraffin chains is also set at 1:2. The weight fraction

of paraffins, naphthenic rings, and aromatic rings are then calculated as below:

WT, = %C, - "^^^'•'•"-'-^^^^•^^^ny-o^'"' (4.20) AW,^,,,„-RATIO,

where

I-paraffin, naphthene ring, and aromatic ring,

C]- weight percentage of carbons of molecule type I in the total carbons in the FCC feed,

wt.%),

RATIOi- the carbon/hydrogen ratio in the molecule type I,

WTi- weight fractions ofthe molecule type I in the FCC feed,

AWcarbon- atomic weight of carbon, 12,

AWhydrogen- atomic weight of hydrogen, 1.

The comparison between the industrial data and the weight percentages ofthe

paraffin, naphthene rings, and aromatic rings in the FCC feed calculated by equation 4.20

are listed in Table 4.3.

The industrial data do not include the volume percentage ofthe light lumps.

However, it can be estimated that the volume percentage of light lumps is between 10

vol.%> and 20 vol.%) since 650°F, the cut point for light lumps, is greater than the highest

possible point for 10 vol.%, 622°F, while less than the lowest possible point for 20 vol.%,

678°F. It can be seen from Table 4.3 that the calculated value, 12.8 vol.%, is in this

range. Since the specific gravity values of paraffin, naphthene, and aromatics in the FCC

feed are unknown, the volumetric percentage of these components can not be calculated

in the feed characterization. However, the calculated weight percentages of paraffin,

naphthene and aromatics are close to the volume percentages of corresponding

components in the industrial data. The wt.% of aromatics is 1.3% higher than its vol.%.

81

Page 97: 5-95 gap chart

which is expected because aromatics have heavier density than other species. The

opposite is true for paraffins.

Table 4.3 Comparison between the Industrial Data and Feed Characterization ofthe FCC Feed.

Feed Property Plant Data, Typical FCCU Model prediction Combined Fresh Feed Analysis, May through August, 1998

Light Components, <650 °F 10% TBP, 606-622 °F 13.4 (vol.%)

20% TBP, 678-694 °F 12.5 (wt.%)

Paraffins 56.9-59.1 (vol.%) 56.1 (wt.%)

Naphthenes 22.2-24.8 (vol.%) 23.8 (wt.%)

Aromatics I8.I-I8.7 (vol.%) 20.0 (wt.%)

The weight percentage ofthe aromatics from the feed characterization, 20.0 wt.%,

is higher than the values ofthe volume percentage ofthe aromatics in industrial data,

18.1-18.7 Vol.%). This is expected because aromatics have higher specific gravity than

those of paraffins and naphthenes with the same boiling point. Hence, the weight

percentage ofthe aromatics in the FCC feed is greater than its volume percentage from

the absolute value. The comparison made here is not rigorous due to the lack of detailed

FCC analysis from industrial data. General speaking, the calculated values from feed

characterization are in good agreement with the industrial data considering that empirical

correlations are used in the calculations.

4.3.3 Weight Fractions ofLumps in the FCC Feed

According to the 10-lump model, the weight percentage of aromatic rings is the

summation ofthe weight percentage of Arh and Ari. Arh and Ari have the same molecular

stmcture. The only difference between the two is their boiling range. This is also tme for

each pair of Ph and Pi, Nh and Ni, and Ash and Asi. An assumption is made that the weight

percentage ofthe heavier lump in the total heavy lumps is equal to the weight percentage

of lighter lump in the same pair in the total light lumps. The assumption is based on the

82

Page 98: 5-95 gap chart

fact that the composition of FCC feed does not change dramatically with respect to its

boiling range. The relative small amount of light lumps, which is about 15 vol.%) ofthe

FCC feed, causes an insignificant error due to this assumption. The weight percentage of

Arh and Ari are calculated using following formula:

WT W,= --!-!- , (4.18)

WT^+WT^+WT/ W,=W„^,,,,^^-W,, (4.19)

^ / . = ^ / - ^ v (4.20)

where

I - aromatic rings, naphthene rings, and peiraffin,

Wi- weight fraction ofthe molecular type I in the FCC feed,

Wii- weight fraction ofthe light lump ofthe molecular type I in the FCC feed,

Wih- weight fraction ofthe heavy lump ofthe molecular type I in the FCC feed,

Wiightiump- weight fraction of total light lumps in the FCC feed.

The weight fractions of lump Arh and lump Ari are equal to WArh and WAFI,

respectively:

%Ar,=W,,^, (4.21)

%An=W,^^, (4.22)

where

%Arh, %)Ari- weight fractions of lump Arh and lump Ari in the FCC feed, respectively.

Only the weight fractions of lump Ath and lump Ari can be directly obtained from

equation 4.18 to 4.22. Since the lump Ash and Itimp Asi can be the paraffin chains or the

naphthenic group, the remaining six lumps belong to either naphthenic molecules or

83

Page 99: 5-95 gap chart

paraffin molecules. No correlation is given in literature to calculate the weight

percentage of aromatic substituent lumps, Ash and Asi.

Ellis (1998) provided the composition of an FCC feed with an API gravity of 23,

which is close to the API gravity ofthe FCC feed, 22.0 to 23.3, studied here. The weight

fractions of eight lumps in the FCC feed of API=23 given by Ellis (1998) are listed in

Table 4.4.

Table 4.4 Weight Fractions of Eight Lumps in an FCC Feed of API=23.

Lump Name Ph Nh Ash Arh Pi Ni Asi

Weight Fraction 0.36 0.15 0.25 0.14 0.04 0.02 0.02

Ari

0.02

Assuming the same ratios of Ash to Arh and Asi to Ari in the feed given in Table

4.4 and the FCC feed studied here, the weight fractions of Ash and Asi are calculated as

below:

o/oAs,=o/oAr,-y'''^'-", (4.23)

o/oAs,=%Arr2.,'^'-'\ (4.24)

where

%Ash, %)Asi- weight fractions of lump Ash and lump Asi in the FCC feed, respectively,

%Arh,APi=23, %)Ari, APi=23, %Ash,APi=23, %Asi,APi=23 -Weight fractions of lumps Arh, Ari,

Arh, and Ari in the FCC feed of API=23, respectively.

The weight fractions of lumps Ph, Pi, Nh, and Ni are calculated using the following

formula:

W %n=(I-fl'„„.,.., -%As, -%ArO-- <^^*^ . (4.25)

paraffin .heavy naphlhene .heavy

84

Page 100: 5-95 gap chart

w %N„ =(1-W„^,,„,^ -o/oAs, -%ArJ-- ".^mUf!^ , (4.26)

paraffin .heavy naphlhene .heavy

W %P, = (W,^„ ,„„, - %As, - %Ar,) - '-^^rsff^ , (4.27)

paraffin,light naphlhene .light

w %N, = (W„^,„„^ - o/oAs, - o/oAr,) - "-^^!^ , (4.28)

paraffin.lighl naphlhene .lighl

where

%Ph, %)Pi ,%Nh, %Ni -weight fractions of lump Ph, lump Pi, lump Nh, lump Ni in the FCC

feed, respectively,

Wparaffm.heavy- Weight fraction of the hcavy paraffin in the FCC feed,

Wparaffinjight- weight fraction ofthe light paraffin in the FCC feed,

Wnaphthene.heavy- Weight fraction of heavy naphthene in the FCC feed,

Wnaphthenejight- weight fraction of light naphthene in the FCC feed.

In order to determine if the feed characterization procedure gives reasonable

results in the operating ranges which are used in the optimization studies, the feed

characterizations were carried out in the operating ranges and the results were compared

with the base case calculated above. The two main operating variables that affect the

composition of FCC feed are the heavy distillate ASTM 95% point and the HVGO TBP

100%) point. The heavy distillate is the adjacent lighter side-draw product ofthe FCC

feed in the atmospheric tower. The heavy distillate ASTM 95% point determines the

amount ofthe light components in the FCC feed. HVGO is the heaviest side-streams

among the four streams composing ofthe FCC feed. The HVGO TBP 100% will affect

the composition of heavy lumps in the FCC feed. The weight fractions of eight lumps and

other properties of FCC feed under different operating conditions are listed in Table 4.5.

85

Page 101: 5-95 gap chart

Table 4.5 Weight Fractions ofthe Eight Lumps in the FCC feed across the Operating Range.

Base Case Case 1 Case 2 Case 3 Case4 Industrial data

Operating conditions

HD ASTM 650

95%, °F

HVGO TBP 1050

100%, °F

630

1050

670

1050

650

1010

650

1090

650

1050

Feed properties

Light lump, vol.% 13.4

Aromatics, wl.%o 20.0

Naphthenes, wt.% 23.9

Paraffins, wt.% 56.1

18.9

19.3

23.0

57.7

7.4

20.7

25.0

54.3

13.7

19.8

25.2

55.0

13.0

20.4

22.5

57.1

10%-20%

18.I-I8.7

(vol.%)

22.2-24.8

(vol.%)

56.9-59.1.

(vol.%)

Weight fraction of lump

Ph 0.272

Nh 0.115

Ash 0.313

Arh 0.175

P, 0.053

Ni 0.022

Asi 0.025

An 0.025

0.272

0.108

0.284

0.159

0.078

0.031

0.034

0.034

0.269

0.124

0.345

0.193

0.028

0.013

0.014

0.014

0.268

0.123

0.308

0.172

0.054

0.025

0.026

0.026

0.272

0.107

0.320

0.179

0.051

0.020

0.025

0.025

In Case 1 and Case 2, the heavy distillate ASTM 95% point is perttirbed by 20°F.

It is observed from Table 4.5 that light lump content increases with the decrease ofthe

heavy distillate ASTM 95%o point and vice versa. It is expected because if the heavy

86

Page 102: 5-95 gap chart

distillate ASTM 95% point drops, more light components are left in the FCC feed. The

opposite is tme when the heavy distillate ASTM 95% point increases.

In Case 3 and Case 4, the HVGO TBP 100% point is perturbed by 40°F. It is

observed from Table 4.5 that the light lump content decreases with the increase ofthe

HVGO TBP 100%) point and vice versa. The change is small compared to the light

components change caused by perturbing the heavy distillate 95% point. This is expected

because the HVGO have a small fraction in the heavy lumps.

4.4 Model Benchmarking

The steady-state model developed by Ellis et al. (1998) is used in the present

study. The steady-state model has been quantitatively benchmarked against published

results. These results show a good representation ofthe available data. However, the FCC

unit in the refinery considered in this work is slightly different from what was described

in the model. The biggest difference is that the FCC unit studied here has only about 70%

of throughput ofthe original model. The feed qualities are also different. The feed quality

significantly affects the product distribution predicted from the FCC model. The FCC

model using the FCC feed characterized above show significant difference in product

distribution than the industrial data. This may be explained by the different catalysts

used. The type of catalyst in the original reaction system that the model is based on is

unknown. Since there are many types of FCC catalysts, it is very likely that catalyst used

in the FCC unit considered in this work is different from catalyst in the unit described by

the original model. In order to represent the FCC unit accurately, ftirther benchmarking

ofthe original model is required.

The original model was benchmarked against the industrial data obtained from the

refinery considered in the work. Only the industrial operating data for one day of

operation is available. The industrial data are listed in Table 4.6.

87

Page 103: 5-95 gap chart

Table 4.6 FCC Model Benchmarking: Industrial Data and Model Prediction.

Process variable

FCC Feed, MLB/D

Reactor temperature, °F

Regenerator temperature, °F

O2 in flue gas, mol%)

FCC gasoline, MLB/D

LCO, MLB/D

HCO, MLB/D

Light Gas, MLB/D

Industrial Data

6378

987

1284

0.011

3433

1096

450

1201

Model Prediction

-

-

3434

1097

454

1206

Relative error (%)

-

-

-

0.03

0.09

0.89

0.42

The most important fiinction ofthe FCC model is to correctly predict the product

distribution. The values ofthe reaction parameters directly affect the product distribution.

Only the activation energy ofthe reactions are used in parameterization because product

distribution is more sensitive to the activation energy than to the frequency factor. Since

only bulk volumes of four products from industrial data are available for the

parameterization, three adjustable factors are used to adjust the values of activation

energy ofthe cracking reactions. Each adjustable parameter corresponds to the formation

of one ofthe three products: FCC gasoline, LCO, and light gas. It can been seen from

Figure 4.2 that HCO is not formed in the FCC cracking. Hence, there is no adjustable

parameter for HCO. Each original value of activation energy is adjusted by one of three

parameters based on the observed products ofthe reaction. For example, the activation

energy ofthe reactions from lump Ph, an component of HCO, to Pi, an component of

LCO, is multiplied by the adjustable parameter of LCO formation and becomes the new

value ofthe activation energy. The adjustable parameters that are used in the

parameterization and their values are listed in Table 4.7.

88

Page 104: 5-95 gap chart

Table 4.7 Adjustable Variables in FCC Model Benchmarking.

Adjustable parameter Description Parameter value

^^gasoline Adj ustable parameter for activation 0.8576

energy of gasoline formation reactions

^LCO Adjustable parameter for activation 0.8507

energy of LCO formation reactions

•Alightgas Adj ustable parameter for activation 1.0198

energy of light gas formation reactions

ki,(lboil)(s)/lb

catalyst

Kinetic frequency factor for the 8.626699E5

formation of gasoline from gas oil

kc,

0.4

Frequency factor for the formation of 0.48

(lb coke/lb catalyst) coke

A3, Btti/lb Ratio of heat transfer coefficient of

preheat box versus gas oil specific

heat

14.33

Hcrack, BTU/(lb oil) Heat for cracking reaction -326.86

Oa.ref, reference O2, in estimating coke on

Mole oxygen/mole air regenerated catalyst

0.01197

ko2, (lbcoke)(s)/ft^ Frequency factor in the rate ofthe

depletion of oxygen

2.55644E9

89

Page 105: 5-95 gap chart

The process constraints are also parameterized to prevent the violations of process

constraints. For this purpose, some process adjustable parameters are also included in the

parameterization. These adjustable parameters have been used to develop the original

model. Their values are also listed in Table 4.7. The values ofthe process constraints are

from the design documents ofthe refinery and the original model. Personnel from the

refinery agreed on the values of these process constraints. The values of process

constraints are listed in Table 4.8. For the details of FCC model, adjustable parameters,

and process constraints, the reader can refer to Ellis (1996).

FCC model parameterization is an optimization problem. The least square type

objective function is used in the parameterization. The objective function is shown

below:

Objfparameterization ~ LI'^I V^model,I ~^base,I) ' (4-29) /

where

Objfparamterization- objective ftuiction of the FCC modcl parameterization,

Wi- weighting factor,

Wmodeij- weight flow rates of product I predicted by the FCC model, MLB/D,

Wbase,!- weight flow rates of product I in industrial data, MLB/D.

NPSOL optimization package (Gill et al., 1986) was used in the parameterization.

Table 4.6 shows the comparison ofthe industrial data and model predictions. For this set

of data, the model predictions fit the industrial data very well. The largest relative error is

less than 1%.

90

Page 106: 5-95 gap chart

Table 4.8 Process Constraints of FCC Unit.

Constraint Lower number Bound

Process Description Variable

Upper Bound

1 500.0

0.0

14.7

Temperature of fresh feed exiting 700

preheat furnace, °F

Flow rate of fuel to preheat 39.5

furnace, SCM

Reactor pressure, psig 30

14.7 Regenerator pressure, psig 30

5100

0.0

-5.0

0.0

sa

r'sucn.lift"

1" surge,! ift

P6-P4

t sp

Actual speed of lift air blower, 6100

RPM

Difference between suction flow None

and surge flow for lift air blower,

ICFM

Reactor-regenerator pressure 2.0

difference, psi

Level of catalyst in standpipe, ft 20.0

10

35000 Fsucn.comb Combustion air blowcr inlet 42000

suction flow, SCFM

0.0 Fwg Flow rate ofwet gas to wet gas 0.67

compressor,

(lbmol)/s

91

Page 107: 5-95 gap chart

4.5 FCC Gasoline Octane Model Modification

Since the FCC gasoline has about 45 vol.% ofthe gasoline pool, the changes of

the properties ofthe FCC gasoline have great effect on gasoline blending. It is highly

desirable to predict the properties of FCC gasoline accurately. In the original model, Ellis

(1996) developed an empirical correlation to related the motor octane number to the riser

temperature and conversion as below:

MON = MON,^, + a, (T, - 7;„,,; + a,(Com - Conv,^,^), (4.30)

where

a i - constants, MON/°F,

a i - constants, MON/conversion

Conv- weight conversion, lb (gasoline+light gas)/ lb gas oil,

Convbase- base weight conversion, lb (gasoline+light gas)/ lb gas oil,

MON- motor octane number ofthe FCC gasoline,

MONbase- motor octane number ofthe base FCC gasoline,

Tf- riser temperature, °F

Tr, base- basc riscr temperature, °F.

The values of base case and constants are listed in Table 4.9.

It is found that MONs calculated from this correlation were about 2 octane

number higher than the corresponding industrial data. In order for this correlation to fit

the industrial data, modification is necessary. One set of data from the industrial data was

selected as the base case. The value ofthe constant al is also changed based on the

industrial data. The new values ofthe base case and constants are listed in Table 4.9.

92

Page 108: 5-95 gap chart

Table 4.9 FCC Octane Model.

Parameter Unit Original Model Current Model

Convbase (gasoline+light gas)/ lb gas oil 0.55 0.80

MONbase, None 72.5 80.7

Tr.base °F 900 995

ai MON/°F 0.05 0.022

a2 MON/conversion 0.17 0.17

93

Page 109: 5-95 gap chart

CHAPTER 5

REFORMER FEED CHARACTERIZATION

AND MODEL BENCHMARKING

5.1 Process Overview

Catalytic reformer is an important unit in gasoline production. The reformate, the

product from the catalytic reformer, is a major blending stock in gasoline pool. Catalytic

reformate furnishes approximately 40% ofthe United States gasoline requirements (Gary

and Handwerk, 1984). The reformer in the refinery considered in this work provides

about 30 vol.% blending stock for the gasoline pool.

To appreciate the importance of catalytic reforming in gasoline production, the

octane number must be understood. Octane number is an index used to measure the

antiknock quality ofthe gasoline. Higher the octane number, the better the antiknock

quality. Octane number is the most important specification for gasoline. Gasoline with

higher octane number normally sells at higher price in the market. Two values of octane

number, research octane number (RON) and motor octane number (MON), are used.

Octane rating is the arithmetic average of MON and RON. Octane rating is used to refer

the general antiknock quality ofthe gasoline in the United States.

The octane ratings ofthe typical gasoline blending stocks in the refinery are

shown in Table 5.1.

Table 5.1 Octane Ratings ofthe Typical Gasoline Blending Stocks.

Gasoline blending stock Octane rating, (R0N+M0N)/2

Low severity reformate 92.3

High severity reformate 94.5

FCC gasoline 87.1

Alkylate 91.1

Light sfraight-run (LSR) gasoline 69.0

Butane 91.5

94

Page 110: 5-95 gap chart

Low severity reformate is the reformate produced when the severity ofthe

catalytic reformer is low and high severity reformate is the reformate produced when the

severity ofthe catalytic reformer is high. The inlet bed temperatures usually indicate the

severity of the operation. The higher the inlet bed temperature, the more severe the

operation. It can been seen from Table 5.1 that the reformate has the highest octane

number among all the blending stock. Considering the large quantity ofthe reformate, it

is clear that the reformate is the major octane booster in gasoline blending.

It is not economical to process the Light straight-mn (LSR) gasoline (C5-180°F)

as feed to the reformer as this fraction is largely composed of low-molecular-weight

paraffins that tend to crack to low-value butane and light fractions (Gary and Handwerk,

1984). The typical feedstock to the catalytic refomer is the heavy naphtha with a boiling

range from 180 to 375°F.

The heavy naphtha produced from the cmde unit has an octane rating around 70,

which is too low for gasoline blending. The reformer converts the heavy naphtha to the

high-octane reformate. Reforming Process are classified as continuous, cyclic, or semi-

regenerative depending upon the frequency of catalyst regeneration. In a continuous

process, the catalyst is removed and replaced during normal operation. In a semi-

regenerative process, the unit is shut down periodically to regenerate the catalyst. The

normal intervals is 3 to 24 months. The cyclic process is a compromise between these

extremes and is characterized by having a swing reactor in addition to those on-stream in

which the catalyst can be regenerated without shutting the unit down (Gary and

Handwerk, 1984). The catalytic reformer in the refinery considered in this work is a

semi-regenerative type. The semi-regenerative type reformer has the advantage of

minimal capital costs compared to continuous and cyclic type reformers (Gary and

Handwerk, 1984). The disadvantage ofthe semi-regenerative type reformer is that the

reformer needs to be taken off-stream periodically to regenerate the reforming catalyst

that looses its activity during the operation.

The reformer plays a key role in the hydrogen usage in the refinery. The reformer

is the main provider ofthe hydrogen that is used in several hydrotreaters in the refinery to

95

Page 111: 5-95 gap chart

treat feeds and final products. Without a reformer, a refinery has to rely on either a steam

reforming unit or purchased hydrogen from resources outside the refinery.

The flow diagram ofthe reformer in the refinery is shown in Figure 5.1. The

reaction system consists of three fixed-bed reactors in sequence. The heavy naphtha from

the cmde unit goes through a naphtha desulfurizer to eliminate the sulfiir and other

impurities in the feed. Then the heavy naphtha is mixed with the hydrogen stream and

exchanges heat with the product stream from the third reactor. The heavy naphtha is

further heated in the preheat fiimace to reach the reaction temperature, about 890 to 980

°F. Since the main reactions taking place in the reactor are endothermic, the temperature

drops sharply in the reactor. The temperature drops to such a level that no reforming

reaction can take place near the exit ofthe reactor. The effluent stream is withdrawn from

the reactor and is heated in an intermediate heater to the reaction temperature. The hot

stream flows to the next reactor.

The product stream exits the third reactor and exchanges heat with incoming

naphtha. The product stream then flows to a separator, which is a flash dmm, to separate

the hydrogen in the product stream. The major part of hydrogen is returned to the reaction

system. The extra hydrogen is used in other units in the refinery. After the separator, the

product stream flows to a stabilizer to separate the C4 and lighter from the heavier

components. C4 and lighter components are fed to the gas plant for ftirther treatment. The

bottom stream ofthe stabilizer is the main product, reformate, which goes to the storage

tank for gasoline blending.

5.2 Reformer Modeling

The reformer modeling has been studied by many researchers. Taskar (1996) gave

an exhaustive literature survey on reformer modeling. The steady-state model developed

by Taskar (1996) was used in this work. The core of Taskar's model is a detailed kinetic

scheme involving 35 pseudo-components connected by a network of 36 reactions in C5-

Cio range using Hougen-Watson-Langmuir-Hinshelwood-type reaction rate expressions.

The compositions of pseudo-components in a typical reformer feed (Lin, 1988; are

listed in Table 5.2.

96

Page 112: 5-95 gap chart

u CO nj QL 0)

• a 0) 0)

>»!: O <!. (U C D) <U C 3 TO £ -S 0 y o <" (J CD (U

a:

,o <+3

ca J5 H-» J3

O H

ca

z o -*-> >^ ca . ^ j

ca U (U

> ca kH 1) c 1) 00 (U

fti • l-H

a (U C/2 ca

l+H

o <D

a <u JS CJ

C/2

^ o

fa c/3 1/3 (U o o kH

fa '—' i n

kH 3

Fig

97

Page 113: 5-95 gap chart

Table 5.2 Chemical Components ofthe Reformer Feed.

Chemical components

Hydrogen

CI

C2

C3

C4

C5-

n-Pentane

Iso-Pentane

Multi-branched hexanes

Single-branched hexanes

n-Hexane

Five-carbon ring C6

naphthenes

Benzene

Multi-branched heptanes

Single-branched heptanes

n-Heptane

Five-carbon ring C7

naphthenes

Six-carbon ring C7

naphthenes

Toluene

Volume fraction of chemical component in a typical reformer feed

0

0

0

0

0

0

0.011533793

0.005766896

0

0.034154728

0.068309457

0.03902341

0.005057543

0.02882307

0.02882307

0.093747762

0.028779382

0.052533793

0.023964035

Molar specific volume. cm^/mole

31.0

52.

68.

84.

101.4

116.1

116.1

117.4

I3I.6

132.9

131.6

113.1

89.4

147.5

147.5

147.5

128.8

128.3

106.8

Molar fraction of chemical component in a typical reformer feed

0

0

0

0

0

0

0.0136791382

0.00676383225

0

0.0353870741

0.071473286

0.0553827156

0.0076970243

0.026907108

0.026907108

0.0875160473

0.0358655286

0.0657239621

0.0305287825

98

Page 114: 5-95 gap chart

Table 5.2 Continued.

Chemical components

n-Octane

Single-branched octanes

Multi-branched octanes

Five-carbon ring C8

naphthenes

Six-carbon ring C8

naphthenes

C8 Eiromatics

n-nonane

Single-branched nonanes

Multi-branched nonanes

Six-carbon C9 naphthenes

C9 aromatics

n-decane

Single-branched decanes

Multi-branched decanes

Six-carbon ring CIO

naphthenes

C10 aromatics

Volume fraction of chemical component in a typical reformer feed

0.123659717

0.030914929

0.030914929

0.028199327

0.056398654

0.050215496

0.055560406

0.013890101

0.013890101

0.080274821

0.03693356

0.01043014

0.00521507

0.00521507

0.024351882

0.013418857

Molar specific volume. cm^/mole

163.5

162

164.4

144.7

143

123.1

179.6

178.2

170.4

159.8

139.6

196

194.5

191.5

176.3

156.8

Molar fraction of chemical component in a typical reformer feed

0.104142811

0.0262767739

0.0258931714

0.0312810878

0.0633059217

0.0555009602

0.0425968892

0.0107328858

0.0112241799

0.0806332576

0.0359961908

0.00732745014

0.00369198002

0.00374981783

0.0221713355

0.0116436799

It is worth noting that some pseudo-components are real chemical species, such as

n-pentane and iso-pentane. The important difference between pseudo-components in the

reformer and lumps used in the FCC model is that all the chemical species in a pseudo-

component ofthe reformer model has the same carbon number while the chemical

99

Page 115: 5-95 gap chart

species of a lump in the FCC model do not have the same carbon number. One reason for

this modeling difference is that the FCC feed is much heavier than the reformer feed. The

FCC feed is so complicated that it is very difficult to use the same approach as the

reformer model. Using detailed pseudo-components in the reformer provides much

required information for the gasoline blending while the feed characterization of reformer

becomes more difficult.

Deactivation ofthe catalyst in the reformer was also modeled. The model was

parameterized by benchmarking against the industrial data obtained from Phillips

Petroleum Company. The reader can refer to Taskar (1996) for detailed discussion ofthe

kinetics ofthe reforming reactions.

5.3 Reformer Feed Characterization

5.3.1 Naphtha Desulfiirizer

In the refinery considered in this work, the reformer feed is composed of heavy

naphtha, a side-draw stream ofthe atmospheric tower and the bottom stream ofthe

naphtha splitter. The reformer feed flows through a naphtha desulfurizer before entering

the reformer. The metals, hydrogen sulfide, ammonia, organic nitrogen and sulfur

compounds will deactivate the catalyst in the reformer and must be removed before the

feed enters the reformer (Gary and Handwerk, 1984). Hydrotreating is used to remove

these impurities from the feed.

The feed to the naphtha desulfurizer includes reformer feed from the cmde unit

and the hydrogen stream from the reformer. The products ofthe naphtha desulfiirizer

consist ofthe stripper overhead gas, desulfurizer separator overhead gas, liquidified

petroleum gas (LPG), and sweetened naphtha. Stripper overhead gas and separator

overhead gas flow to the gas plant for further treatment. The LPG flows to the

debutanizer in the cmde unit.

A detailed model was not built for the naphtha desulfurizer. Instead, material

balance and fixed ratios are used to calculate the amotmts ofthe hydrogen, the stripper

overhead gas, the separator overhead gas, LPG, and the sweetened reformer feed. The

100

Page 116: 5-95 gap chart

ratios of each operation mode. Summer Mode or Winter Mode, are based on the

industrial data of a typical day in each operation mode. The ratios are listed in Table 5.3.

Table 5.3 Ratios in the Calculation ofthe Naphtha Hydrotreater.

Rates Summer Mode Winter Mode

H2/Feed, MLB H2/MLB feed

H2/Feed, BBL H2/BBL feed

Reformer feed/total feed,

MLB reformer feed/MLB total feed

Reformer feed/total feed,

BBL reformer feed/BBL total feed

Overhead gas/total feed,

MLB overhead gas/MLB total feed

Overhead gas/total feed,

BBL overhead gas/BBL total feed

LPG/total feed,

MLB LPG/MLB total feed

LPG/total feed,

BBL LPG/BBL total feed

0.06997

0.07756

0.9108

0.06366

0.07553

0.8920

0.9020

0.0029

0.0024

0.0509

0.0623

0.9035

0.0017

0.0016

0.0183

0.0255

The amount ofthe hydrogen required by the naphtha desulfurizer is calculated

using the formulas given below:

^H2 = ^reformerfeed ' ^TIO^j21 reformer feed, W '

^Hl = Reformer feed " ^^10^^21 reformer feed, V '

(5.1)

(5.2)

where

WH2- mass flow rate ofthe hydrogen used in the naphtha desulfurizer, MLB/D,

VH2- volumetric flow rate ofthe hydrogen used in the naphtha desulfurizer, BBL/D,

101

Page 117: 5-95 gap chart

Wreformer feed" niass flow rate of the reformer feed from the crude unit, BBL/D,

Vreformer feed- volumctric flow rate of the reformer feed from the cmde unit, BBL/D,

RatioH2/reformer feed, w" niass ratio between the hydrogen and the reformer feed from the

crude unit,

RatioH2/reformer feed, v" volumctric ratio between the hydrogen and the reformer feed from

the crude unit.

The volume and the weight ofthe total feed are the summations ofthe volumes

and the weights of incoming streams, reformer feed and hydrogen. For each product

stream, the volume and the weight are calculated by the formulas given below:

W,=W,,,,-RATIO,„, (5.3)

V,=V^^^,-RATIO^,., (5.4)

where

i- stripper overhead gas, separator overhead gas, LPG, sweetened reformer feed,

RATIOi,v- volumetric ratio between the product i and the total feed to the naphtha

desulfiirizer,

RATIOi,w- mass ratio between the product i and the total feed to the naphtha

desulfurizer,

Vi- volumetric flow rate ofthe product stream i, BBL/D,

Vfeed- volumetric flow rate of total feed to the naphtha desulfurizer, BBL/D,

Wi- mass flow rate of product stream i, MLB/D,

Wfeed- mass flow rate of total feed to the naphtha desulfurizer, MLB/D.

The volumefric flow rate and mass flow rate ofthe sweetened reformer feed

calculated above are then used as the inputs to the reformer model.

102

Page 118: 5-95 gap chart

5.3.2 Paraffins-Naphthenes-Aromatics (PNA) in the Reformer Feed

The molar flow rates ofthe pseudo-components in the feed are required by the

reformer model. The feed characterization is to calculate the molar flow rates of these

pseudo-components in the feed.

It can been seen from the FCC feed characterization that feed characterization is a

complicated process. For a much more complicated reformer pseudo-component system,

no simple empirical correlation is available to calculate the molar flow rates. Taskar

(1996) divided the reformer feed into several cuts based on boiling range and used

empirical correlations (Daubert, 1994) to calculate the physical properties of each cut.

Using the boiling point and the physical properties ofthe pseudo-components in the feed,

the molar percentages ofthe pseudo-components are calculated by using Nelder-Mead

optimization algorithm (Riggs, 1994) to find the composition which will give the same

physical properties of each cut ofthe feed.

One disadvantage of this approach is that the boiling point and the physical

properties of each cut are not accurate because ofthe inaccuracy ofthe empirical

correlations. Another disadvantage is that there may be several combinations of pseudo-

components which can give the same physical properties. In addition, local minimums

exist in the optimization searching which means that the unique solution is not

guaranteed in every optimization searching. If this approach were used in the refinery-

wide model, it will cause erroneous results in the reformer model when the optimization

routine gives unreasonable reformer feed compositions. Although the paraffin-

naphthenes-aromatics (PNA) calculated from the pseudo-component molar flow rates are

close to the industrial data, the molar flow rates of pseudo-components calculated using

the approach above can not be verified due to the lack of industrial data.

The attempts to use the above approach failed in this work. The optimization

algorithm did not give a reasonable solution. This may be due to the inaccuracy ofthe

empirical correlations used to calculate the physical properties. However, some data from

laboratory analysis ofthe portion ofthe cmde oil in the reformer feed boiling range are

103

Page 119: 5-95 gap chart

available in the industrial data. It is more accurate to calculate the reformer feed

information from these laboratory analysis than from empirical correlations. The

industrial data are listed in Table 5.4 through Table 5.7.

Table 5.4 Industrial Data of Crude A.

Property

Yield

API Gravity

Aromatics

Naphthenes+2x

Aromatics (vol.%))

RON

MON

Sulfiir

Unit

Vol.%

API

Vol.%

Vol.%

None

None

wt.%)

Light naphtha I50-200°F

2.5

67.69

1.515

37.93

71.832

67.992

0.007

Heavy naphtha 200-360°F

16.0

55.597

5.692

54.195

49.193

48.730

0.009

Extra heavy naphtha 360-385°F

3.1

46.688

9.474

74.459

33.907

34.407

0.014

Table 5.5 Industrial Data of Cmde B.

Property

Yield

API Gravity

Aromatics

Naphthenes+2x

Aromatics, (vol.%))

RON

MON

Sulftir

Unit

Vol.%

API

Vol.%

Vol.%

None

None

Wt.%

Light naphtha 150-200°F

1.9

73.961

8.556

40.037

71.832

67.992

0.013

Heavy naphtha 200-360°F

9.7

55.534

18.169

60.471

42.657

41.863

0.028

Extra heavy naphtha 360-385°F

2.8

44.077

23.499

78.858

34.648

34.408

0.037

104

Page 120: 5-95 gap chart

Table 5.6 Industrial Data of Crude C.

Property

Yield

API Gravity

Paraffins

Naphthenes

Aromatics

RON

MON

Sulfiir

Unit

Vol.%

API

Vol.%

Vol.%

Vol.%

None

None

Wt.%

Light naphtha 155-265°F

10.33

63.1

62.3

31.9

5.8

-

-

<.02

Heavy naphtha 265-350°F

9.35

51.7

55.4

28.5

16.1

-

-

0.03

Extra heavy naphtha 350-400°F

4.86

45.8

42.4

42.6

15.0

-

0.05

5.7 Industrial Data of Cmde D.

Property

Yield

API Gravity

Paraffins

Naphthenes

Aromatics

RON

MON

Sulftir

Unit

Vol.%

API

Vol.%

Vol.%

Vol.%

None

None

Wt.%

Light naphtha 155-265°F

9.18

60.9

57.4

34.3

8.3

-

-

0.02

Heavy naphtha 265-350°F

9.78

51.7

55.8

27.6

16.7

-

0.02

Extra heavy naphtha 350-400°F

5.44

46.1

41.8

45.8

12.4

-

-

0.02

The naphtha desulfurizer that treats the reformer feed mainly removes the sulfur

from it. Hydrocarbon in the reformer feed does not react in the naphtha desulfurizer.

Hence, the PNA ofthe sweetened reformer feed after the naphtha desulfurizer is

considered to be the same as the reformer feed entering the naphtha desulfurizer. The

105

Page 121: 5-95 gap chart

normal boiling range ofthe reformer feed is 180 to 375°F. It can be seen from Table 5.4

through Table 5.7 that all or part of light naphtha, heavy naphtha, and extra heavy

naphtha is in the boiling range ofthe reformer feed. Therefore, the PNA ofthe reformer

feed can be calculated from the PNAs of these cmde cuts.

The cmde cut ofthe light naphtha is the source for both the light straight-mn

(LSR) gasoline and the reformer feed. The portion of light naphtha in the reformer feed is

calculated from the volume corresponding to the EBP ofthe light naphtha and the volume

corresponding to the cut point ofthe LSR, which has been calculated in the naphtha

splitter model in Chapter 3. The calculation uses the follow formula:

V -V n .,. '^ EBP.In ' cul poinl. LSR / c ^N

Portion^^j„„„^^j^^j,„= —^ , (5.5)

where

PortioUreformer feed. In- volumetric portion of the light naphtha in the reformer feed,

VEBP.III- volume percentage corresponding to the EBP ofthe light naphtha on the whole

crude TBP- volumetric curve, vol.%,

Vcut point, LSR- volumc percentage corresponding to the cut point ofthe light naphtha on

the whole crude TBP- volumetric curve, vol.%),

Vin- volume percentage the light naphtha in the total cmde oil, vol.%.

It is to be noted that the EBP ofthe light naphtha ofthe cmde A and cmde B is 200°F

while the EBP of cmde C and cmde D is 265°F.

It can been seen from Table 5.4 and Table 5.5 that the industrial data ofthe cmde

A and cmde B have the volumetric percentages ofthe aromatics and

naphthenes+2x Aromatics. The volumetric percentages ofthe naphthenes and paraffins

can be calculated by the following equations:

%naphthene = %N2A - %aromatics, (5.6)

%paraffin = 100 - %aromatics - %naphthene, (5.7)

106

Page 122: 5-95 gap chart

where

%)paraffin, %naphthene, %aromatics- the volume percentages of paraffin, naphthene,

aromatics, respectively, vol.%,

%)N2A- the volume percentages of naphthene+2xaromatics, vol.%).

Now we have the PNA information of all the cmde types in the cmde feed. The

PNA ofthe reformer can be calculated as the summation ofthe PNAs ofthe cmde cuts in

the reformer feed:

(Portion -V -V +\ 0/ ^ V^O/T/ ^^^^ 'reformer feed.ln.i '^ In.i '^X.ln.i^ A)Ji = / /OV: -

^-' V -V +V -V I V hn.i ' X.hn.i^ ' ehn.i " X.ehn.i

(5.7)

where

i- cmde type A, B, C, and D,

X- molecule type, paraffin, naphthene, and aromatics,

%)X- volume percentage of molecule type X, paraffin, naphthene, aromatics, vol.%,

%)Vi- volume percentage of cmde type i in the total cmde, vol. %,

Vin,i- volume fraction ofthe light naphtha in the cmde type i,

Vx,in,i- volume fraction ofthe molecule type X in the light naphtha in the cmde type i,

Vhn,i- volume fraction ofthe heavy naphtha in the cmde type i,

Vx,hn,i- volume fraction ofthe molecule type X in the heavy naphtha in the cmde type i,

Vehn,i- volume fraction ofthe extra heavy naphtha in the cmde type i,

Vx,ehn,i- volumc fraction ofthe molecule type X in the extra heavy naphtha in the cmde

type i.

The typical PNA ofthe reformer feed is available in the industrial data. The

comparison between the PNA calculated above and the typical industrial data of the

refinery considered in this work are shown in Table 5.8. The operating variables ofthe

cmde unit use the normal values in Table 3.7 in Chapter 3. It can be seen from Table 5.5

107

Page 123: 5-95 gap chart

53

33

14

_

0.1

1.9

1.8

1.3

that the average absolute difference between the results ofthe feed characterization and

typical industrial data is 1.3 vol.%. The biggest difference exists in naphthene with 1.9%.

Table 5.8 Comparison of Volume Percentages of PNA from Feed Characterization and the Industrial Data.

Molecule Type Feed Characterization Industrial Data Difference, vol.%

Paraffin, vol.% 5Z9

Naphthene, vol.% 34.9

Aromatics, vol.% 12.2

Average difference -

5.3.3 Molar Flow Rates ofthe Pseudo-Components in the Reformer Feed

The calculation ofthe molar flow rates ofthe thirty-five pseudo-components in

the reformer feed follows the steps given below:

a. Constmct a typical reformer feed. The volume fractions ofthe pseudo-components in

the typical reformer feed are obtained from literature (Lin, 1988; Turpin 1994;

Taskar, 1996).

b. Assume the ratios ofthe pseudo-components in each molecule type, the paraffins, the

naphthenes, and the aromatics, do not change. Calculate the volume fractions ofthe

pseudo-components in the reformer feed using the formula given below:

p ^ . . = F , . . ^ ^ ^ ^ , (5.8)

where

j - paraffins, naphthenes, and aromatics,

Vi j - volume fraction of a pseudo-component i which is a molecule type j ,

Vij,typicai- volume fraction of a pseudo-component I which is a molecule type j in the

typical reformer feed.

108

Page 124: 5-95 gap chart

Vj- volume fraction of a molecule type j ,

Vj.typicai- volume fraction of a molecule type j in the typical reformer feed,

c. Calculate the molar flow rate of each pseudocomponent using the molar specific

volume ofthe pseudo-component and its volumetric flow rate using the formula given

below:

/ w o / , . = — ^ , (5.9) " Vmol^j

where

molij- molar fraction of a pseudo-component i which is a molecule type j ,

Vmolij- molar specific volume of a pseudo-component i which is a molecule type j .

The volume fractions ofthe pseudo-components are listed in Table 5.2. The molar

specific volumes of pseudo-components are also listed in Table 5.2. The volume fractions

and molar fractions of pseudocomponents calculated in equations 5.8 and 5.9 are

normalized after all the pseudo-components have been calculated.

The total molar flow rate ofthe reformer feed is calculated following the steps as

given below:

a. Calculate the mass flow rate of each cmde cut using the API gravity ofthe cmde cut,

the volume percentage with respect to the particular cmde, and the volumetric flow

rate ofthe particular cmde.

b. Calculate the molecular weight ofthe each cmde cut using the approach described in

the Chapter 3 of this dissertation.

c. Calculate the molar flow rate of each cmde cut using the molecular weight and the

mass flow rate ofthe cmde cut.

d. Calculate the molar flow rate of reformer feed by summing the molar flow rates of

the individual crude cuts.

109

Page 125: 5-95 gap chart

The molar flow rate of a pseudo-component in the feed is then calculated as

below:

Fmol,j=mol,j-V. -Fmol^^^,, (5.10)

where

Fmolij- molar flow rate of a pseudo-component i which is a molecule type j , mol/s,

Fmolfeed- total molar flow rate ofthe reformer feed, mol/s.

The calculated molar flow rates ofthe pseudo-components in the reformer feed

then become the input to the reformer model.

5.4 Reformer Model Benchmarking

The reformer model developed by Taskar (1996) is based on a reformer in

Phillips Petroleum Company. The reformer in the refinery considered in this work is

slightly different from the reformer in Phillips Petroleum Company. For example, the

throughput and catalyst load ofthe reformer considered in this work are significantly less

than those in the Phillips's unit. Some minor modifications in the model have been made

to enable the reformer model to represent the reformer considered in this work.

The amounts of catalyst in each bed ofthe reformer unit considered in this work

have been obtained from the industrial data. They are shown in Table 5.9.

Table 5.9 Catalyst Weight in Each Reactor Bed.

Bed number

1

2

3

4

Total

Catalyst weight original Model,

25000

25000

25000

50000

125000

in lb.

the Industrial Data, lb,

10781

21965

44652

0

77398

110

Page 126: 5-95 gap chart

It can be seen from Table 5.9 that the reformer in the refinery considered in this

work loads has about 40% less catalyst in the reactor than the unit described by the

Taskar (1996). In addition, there are four beds in the Phillip's unit while only three beds

are used in the reformer considered in this work.

The reaction system also needs to be modified to fit the industrial data. The

reformer operation in the refinery considered here has two operation modes: low severity

and high severity. The typical operation data ofthe low-severity operation mode and the

high-severity operation mode were used in the model benchmarking. The industrial data

are listed in Table 5.10.

Table 5.10 Comparison ofthe Industrial Data and Model Prediction after Benchmarking.

Operation

Mode

Low severity Conversion

Research Octane number

Aromatics in the reformate.

vol.%

Hydrogen Production,

SCF H2/BBL feed

Industrial

data

0.802

95.

59

1176.54

Model

prediction

0.797

94.9

55

1011.74

Relative

error, %

0.62

0.11

6.78

14.01

High severity

Benzene in the reformate, wt. % 2.7

Conversion 0.757

Research Octane number 99.7

Aromatics in the reformate,

vol.% 63

Hydrogen Production,

SCF H2/BBL feed 1260

Benzene in the reformate, wt. % 3.5

Average Error

1011.74

2.5

0.756

100.

62.4

1150.43

3.45

-

14.01

7.41

0.13

0.30

0.95

8.70

1.43

4.04

111

Page 127: 5-95 gap chart

The activation energy and frequency factor ofthe thirty-six reactions were

considered as adjustable parameters to benchmark the model against industrial data. All

these thirty-six reactions belong to four reaction types: hydrocracking, ring closure,

dehydrogenation, and isomerization. The reader can refer to Taskar (1996) for detailed

description ofthe reaction system. It is found that the model predictions are not sensitive

to the activation energy and frequency factor ofthe isomerization reactions. Therefore,

only the activation energy and the frequency factor ofthe reactions of hydrocracking,

ring closure, dehydrogenation are adjusted in the benchmarking. In order to simplify the

parameterization, two adjustable parameters are used to adjust the activation energy and

frequency factor of all the reactions belong to one reaction type. Therefore, there are total

six adjustable parameters were used in the benchmarking. These parameters adjust the

activation energy and the frequency factor ofthe reactions by multiplying the original

values to form the new values. The values ofthe adjustable parameters that were obtained

in the benchmarking are listed in Table 5.11.

Table 5.11 Adjustable Parameters in the Reformer Model Benchmarking.

Adjustable parameter Description Parameter value

^hydrocracking

r hydrocracking

Aring closure

F ring closure

Adehydrogenation

F dehydrogenation

Adjustable parameter for the activation

energy of hydrocracking reactions

Adjustable parameter for the frequency

factor of hydrocracking reactions

Adjustable parameter for the activation

energy of ring closure reactions

Adjustable parameter for the frequency

factor of ring closure reactions

Adjustable parameter for the activation

energy of dehydrogenation reactions

Adjustable parameter for the frequency

factor of dehydrogenation reactions

1.7996

1.000248EI7

0.6993

5.004975E-6

l.OOI

1.000261

112

Page 128: 5-95 gap chart

The parameterization is an optimization problem. The optimization routines try to

find a set of values ofthe adjustable variables to minimize the difference between the

model predictions and the corresponding industrial data. Like the FCC model

benchmarking, a least square error type function is used as the objective function for the

optimization problem. The formula is given below:

^bjJparamelerizalion ~ 2-1 '^ "'"del.! ~-^hase.I y ' ( • ' • i ^ J

/

where

I - operation variables, such as conversion, RON, etc.,

Objfparamterization- objcctive functiou valuc of the reformer model parameterization,

Wj- weighting factor,

Xmodei,!- valuc of the operation variable I predicted by the reformer model,

Xbase,!- value ofthe operation variable I in industrial data.

NPSOL optimization package (Gill et al., 1986) was used in the parameterization.

Table 5.10 shows the comparison ofthe industrial data and model predictions. The

average relative error ofthe model predictions compared to the industrial data is 4.04%.

The largest error exists in hydrogen production with 14.01% in Summer Mode and 8.70%

in Winter Mode. Since the refinery produces more hydrogen than what it needs, the extra

hydrogen is sent to the fuel gas. Therefore, the error on the prediction of the hydrogen

production has insignificant effect on the economy ofthe refinery-wide operation. The

difficulty in the reformer model benchmarking is that one set of adjustable parameters

was used to benchmark against the data of both operation modes. Using two sets of

adjustable variables for two operation modes may make the model predictions agree with

the industrial data better. However, it is incorrect theoretically to use two sets of

adjustable parameters. The reason is that the basic kinetics are the same for both the high

severity mode and the low severity mode. Hence, the same reaction constants, activation

energy and frequency factor, should be used for both modes. To conclude, the model

113

Page 129: 5-95 gap chart

predictions reasonably agree with the industrial data after benchmarking using one set of

adjustable parameters.

5.5 Average Reformer Operation

The reformer is operated at low severity mode and high severity mode

altematively. Low severity mode uses lower inlet bed temperature which produces

reformate with research octane number (RON) around 95. High severity mode uses

higher inlet bed temperature which produces reformate with RON around 100. Which

severity the reformer is operated is decided by gasoline blending situation. When the

average octane number ofthe whole gasoline pool is high, the reformer is operated at low

severity. On the contrary, the reformer is operated at high severity when the average

octane number ofthe whole gasoline pool is low.

The reformer is also not operated continuously. The reformer is shut down

periodically to regenerate the catalyst which loses it activity during the operation. The

regeneration time is about 2 weeks. The regeneration costs about $450,000 each time.

The reformer is not operated at the same operation mode and is not operated

continuously. However, the refinery-wide model of which the reformer is one unit is a

steady-state model. Hence, the operation shifting and periodical shut down ofthe

reformer must be accounted.

5.5.1 Operation Time Fraction of Low Severity

In order to represent the fact that the reformer altemates operation mode during

each reformer operation cycle, two executions ofthe reformer are carried out in one

execution ofthe refinery-wide model. One execution is for the high severity mode and

the other execution is for the low severity mode. In order to represent the fact that

different mode is used for different time, the product rates from each execution of

reformer model are multiplied by the time fraction discussed above. According to the

engineers from the refinery, the reformer is operated at low severity about 70%) ofthe

total operation time. This operation time fraction of low severity becomes a handle for

adjusting reformer operation. Hence, this fraction is a decision variable for the refinery-

114

Page 130: 5-95 gap chart

wide optimization. The upper limit ofthe operation time fraction of low severity is set at

0.8 and the lower limit is set at zero according to engineers in the refinery.

5.5.2 On-Stream Factor

The On-stream factor is the fraction of time that the reformer is at operation in the

total time. The cycle length ofthe reformer can be approximately calculated using the

formula given below:

C rn _ coke .final /c i '^\ ^cycle - - ^ ' P - i ^ ;

coke .average

where

Tcycie- cycle length, hr,

Ccoke,fmai- cokc contcut at the end of a cycle, kg coke/kg catalyst,

ACcoke,average- the average coking rate in a cycle.

A cycle usually ends when the octane number ofthe reformate can not be

maintained at a certain level even using the highest inlet bed temperature and maximum

H2 recycle ratio. Since octane requirement ofthe reformate for the low severity mode is

different from that ofthe high severity mode, the corresponding coke contents are

different. In this study. Lower limit of octane requirement for each operation mode, the

highest inlet bed temperature and maximum H2/Hydrocarbon recycle ratio are listed in

Table 5.12.

115

Page 131: 5-95 gap chart

Table 5.12 Reformer Operation Limit.

Low severity High severity

Research octane number (RON) 95

Highest Inlet bed temperature, °F 980

Maximum H2/Hydrocarbon recycle

ratio, mol H2/mol Hydrocarbon 7.5

99.5

980

7.5

Ccoke.finai is the cokc contcut at the end of a cycle defined by these operating

conditions. It is known the coke contents on catalysts are not the same across the three

reactor. There are three reactors in the reformer unit. Each reactor is divided into several

regions according to coke content and the coke content in each region is assumed to be

the same (Taskar, 1996). An example ofthe coke contents in different regions is shown

in Table 5.13. The coke contents shown in Table 5.13 are the coke contents at the end of

the cycle ofthe low severity mode.

Table 5.13 Coke Contents in Dfferent Regions.

Section number Low severity, coke content, kg High severity, coke content,

coke/kg catalyst kg coke/kg catalyst

la 3.31 E-2

lb 8.02 E-3

Ic 3.90 E-3

2a 6.36 E-2

2b 8.42 E-2

3a 9.07 E-2

3b 1.02 E-1

(la+2a+2b+3a+3b)/5 7.47 E-2

2.41 E-2

5.76 E-3

2.91 E-3

4.58 E-2

6.21 E-2

6.71 E-2

7.55 E-2

5.49 E-2

116

Page 132: 5-95 gap chart

The first digit ofthe section number indicate which reactor the section belongs to.

The second digit indicates the location of section in the reactor. For example, section 1 b

is the section behind section la and section Ic is the section behind section lb, all in

reactor 1. It can be observed from Table 5.13 that coke contents in section la, 2a, 2b, 3a,

3b have the same order and are much higher than the coke contents in section lb and Ic.

The phenomena have been explained in detail in Taskar (1996). To simplify the

calculation, an average coke content is calculated using the coke contents in section 1 a,

2a, 2b, 3a, and 3b. This average coke content is used as the Ccoke.fmai.

The steady-state model ofthe reformer used in the refinery-wide model has zero

coke content on the catalyst, which represents the catalyst condition at the beginning of

each cycle. The coking rate, with an unit of kg coke/kg catalyst/hr, under this condition is

calculated in the model. However, equation 5.12 requires the value of average coking rate

in one cycle. The relations between the coking rates at zero coke content and the average

coking rates in one cycle was also obtained from the above tests and are listed in Table

5.14. Again, the average coking rate of section la, 2a, 2b, 3a, 3b is used for both coking

rates. The average coking rate in one cycle is calculated as the product of the coking rate

at zero coke content and coking rate ratio for either the low severity mode or the high

severity mode.

Table 5.14 Ratios between Aerage Coking Rate and Coking Rate at the Beginning of a Cycle.

Section number

la

2a

2b

3a

3b

(la+2a+2b+3a+3b)/5

Coking rate ratio

Low severity

4.19

5.19

12.2

4.85

8.25

6.94

Coking rate ratio

High severity

1.92

2.10

3.22

2.01

2.61

2.37

117

Page 133: 5-95 gap chart

After the cycle length is calculated, the on-stream factor is calculated using the

formula given below:

F , = ^-^^ , (5.13) on-slream T T

cycle regeneralion

where

Tregeneration- reformer shut down time or catalyst regeneration time, hr.

According to engineers in the refinery considered in this work, the regeneration

time is about two weeks. Hence, the total hours in two weeks, 336, is used for Tregeneration

for both low severity mode and high severity mode.

In order to account for periodically shut down ofthe reformer, the product rates

from the reformer model are discounted by multiplying each product rate by the on-

stream factor and operation time fraction using the equation given below:

V =v - F T rs M"* product produci .model on-slream operalion lime fraclion ' v*- '-*- ' /

where

Vproduct- volumetric flow rate of a product from the reformer as an output to other units,

bbl/day,

Vproduct, model- volumctric flow rate of a product calculated from the reformer model,

bbl/day,

Toperation time fraction" Operation time fraction of either the low severity mode or the high

severity mode.

5.5.3 Regeneration Cost

The cost of each regeneration ofthe reformer catalyst is normally $450,000.

Refinery-wide optimization requires the regeneration cost for each day. For each

118

Page 134: 5-95 gap chart

operation mode, low severity and high severity, the regeneration cost for each day is

calculated by dividing the total cost of each regeneration by the cycle length calculated in

equation 5.12. The regeneration cost for each day for the reformer as a whole is then

calculated using the equation given below:

regeneralion.reformer regeneralion Jow severily low severity ,^ ^ C\

regeneralion.high .severity \ low severily ^

where

L-OStregeneration,reformer, C^OStregeneration,low severity, C^OStregeneration,high severity— rCgncrat lOn COStS 01

the reformer as a whole, low severity, high severity, respectively,

Tiow severity- Operation time fraction ofthe low severity mode.

119

Page 135: 5-95 gap chart

CHAPTER 6

MODELING OF GAS PLANT,

ALKYLATION UNIT AND

DIESEL HYDROTREATER

Besides the models ofthe four main processing units presented in the previous

chapters, there are other processing units in the refinery, whose models are required to

carry out plant-wide optimization studies. The models ofthe gas plant, the alkylation

unit, and the diesel hydrotreater are discussed here.

6.1 Gas Plant

The gas plant processes all the light gases generated by various units in the

refinery. The main sources of light gases are the cmde unit, the FCC unit, and the

reformer. The naphtha desulfurizer and diesel hydrotreater also produce some light gases

as by-products. All the light gases are combined together and sent to the gas plant. The

product streams coming out ofthe gas plant include fiiel gas, propylene, and alkylation

feed. The fuel gas is used as fiiel in the furnaces in the refinery. Propylene is sold directly

in the market. The alkylation feed stream flows to the alkylation unit to make alkylate, a

gasoline blending stock.

Simplified models were built for the units in the gas plant based on material

balances. The gas plant is very complicated in the sense that the light gas streams come

from multiple sources and the streams are mixed in several places. It is very difficult to

calculate the flow rates ofthe light gas components in all the streams in the gas plant. To

simplify the modeling, the gas plant is divided into two blocks, fuel gas production and

depropanizer. Only the compositions ofthe streams around these two blocks are

calculated.

All the light gas streams from other units in the refinery enter the fuel gas

production block. The fuel gas and the heavier hydrocarbon are two streams exiting the

fuel gas production block. The heavier hydrocarbon stream goes to the depropanizer.

Propylene, known as C3 product in the refinery considered in this work, exits from the

120

Page 136: 5-95 gap chart

top ofthe depropanizer and is sold as a final product. The stream that exits from the

bottom ofthe depropanizer becomes the feed stream to the alkylation unit. The

schematic ofthe gas plant is shown in Figure 6.1.

6.1.1 Sources of Light Gas

Light gas mainly comes from the cmde unit, the FCC unit, and the reformer unit.

The light gas produced in the cmde unit was calculated based on the LP reports obtained

from the refinery considered in this work. The light gases from the cmde unit are listed in

Table 6.1. The same product rates are used to calculate the amounts of light gases from

the cmde unit in the model disregarding the operating conditions in the cmde unit.

Light gas from crude unit

Light gas from FCC

Light gas from reformer

i

Fuel gas

Fuel gas production block

C3/C4 stream

C3 product

Depropanizer

Alicylation feed

Figure 6.1 Schematic ofthe Gas Plant in the Fuel-Oriented Refinery.

121

Page 137: 5-95 gap chart

Table 6.1 Light Gas Production Rates in the Cmde Unit.

Light gas Summer Mode Winter Mode Difference

Ethane, lb/lb cmde feed 0

Propane, lb/lb crude feed 0.00175

Isobutane, lb/lb cmde feed 0.00202

Normal Butane,

Lb/lb crude feed 0.00593

0

0.00054

0.00217

0

0.00121

0.00015

0.00625 0.00032

The largest amount of light gas produced in a refinery comes from the FCC unit.

In the original FCC model of Ellis (1996), the light gases are calculated by the empirical

correlations based on the API gravity ofthe FCC feed and the conversion ofthe FCC

feed. However, the predictions from the empirical correlations are not consistent with the

industrial data obtained from the refinery considered in this work. Hence, the industrial

data for light gas production rates in the FCC unit are used here. The percentages ofthe

light gas components in the total light gas ofthe FCC unit are shown in Table 6.2.

It can be seen from Table 6.2 that both modes have similar light gas production

exception propane. This is because the operation ofthe FCC unit in the Winter Mode is

almost the same as that ofthe Summer Mode. The average values listed in Table 6.2 are

used to represent the components of light gases modeled for both operation modes. The

FCC model can only predict the total amount ofthe light gas, but not individual

components present in the light gas. Assuming the same weight percentages ofthe light

gas components in the total light gas as those in the industrial data, the weights ofthe

light gas components from the FCC unit are calculated using the formula given below:

^ti = ^ttotal -^i^ (6.1)

Where

wti- the weight ofthe light gas component i,

v^totai- the total weight ofthe light gas.

122

Page 138: 5-95 gap chart

Xi- the weight fraction of light gas component i in the total light gas.

Table 6.2 Compositions of Light Gas from FCC Unit.

Light gas

Gases < C2

Propane

Propylene

Isobutane

Butylene

Normal Butane

Summer Mode

wt.%)

9.3

29.3

17.1

21.7

9.0

13.6

Winter Mode

wt.%

9.5

29.2

16.9

21.6

9.1

13.6

Average

wt.%)

9.4

29.2

17.0

21.6

9.1

13.6

The light gas components from the reformer have already been calculated by the

detailed reformer model discussed in Chapter 5 of this dissertation. However, the

reformer model only predicts the total amount of butane, which includes both normal

butane and isobutane. Therefore, it is assumed that isobutane is about 42 wt.% ofthe

total butane according to the industrial data obtained from the refinery considered in this

work.

6.1.2 Fuel Gas Production

The fuel gas mainly contains the hydrocarbons with carbon number less than three

and hydrogen. The light gas stream from the reformer contains some hydrogen. The fuel

gas also contains a small amount of heavy hydrocarbons with carbon number equal to or

greater than three. These heavier hydrocarbon components enter the fuel gas due to

incomplete separation in the gas plant. Since we do not have rigorous models for each

separation units in the gas plant, the average industrial data ofthe loss of heavier

hydrocarbons in the fuel gas shown in Table 6.3 are used in the model. These losses are

assumed to be constants for each operation mode and do not change with operating

conditions. The total amotmt ofthe fuel gas is the sum ofthe light hydrocarbons and the

losses of heavy hydrocarbons in the fuel gas.

123

Page 139: 5-95 gap chart

Table 6.3 Losses of Heavy Hydrocarbons in the Fuel Gas.

Light gas

Propane, wt.%

Propylene, wt.%o

Isobutane, wt.%

Butylene, wt.%)

Normal Butane, wt.%)

Summer Mode

33.4

26.3

16.9

6.2

11.4

Winter Mode

32.0

26.1

16.8

6.2

8.2

6.1.3 Depropanizer

The depropanizer separates the heavier hydrocarbon stream into C3 product and

alkylation feed. The C3 product, the overhead stream, contains propylene and propane.

The alkylation feed, the bottom stream, contains mainly normal butane and isobutane. It

also contains some propylene and propane due to incomplete separation in the

depropanizer. Propylene can also react in the alkylation unit to form alkylate. From the

industrial data obtained from the refinery considered in this work, the C3 product stream

does not contain any C4 hydrocarbons while the alkylation feed stream contains some C3

hydrocarbons. If the distributions ofthe C3 hydrocarbons in the two streams are known,

the compositions ofthe two streams can be calculated. The industrial data for propylene

entering the C3 product stream shown which are in Table 6.4 are used in the model.

These values do not change with respect to operating conditions.

Table 6.4 Portion ofthe Propylene in the C3 Product.

Mode

Portion of total propylene entering C3 product, wrt.%

Summer

Mode

60.7

Winter Mode

79.4

The mass flow rate ofthe propylene in the C3 product can then be calculated from

the total mass flow rate ofthe propylene in the depropanizer feed and the values in Table

6.4. There is no industrial data about the portion of propane entering the C3 product

124

Page 140: 5-95 gap chart

stream available. However, it is known from the industrial data that the propane is about

40 vol.%) and the propylene is about 60 vol.%) in the C3 product. This value is valid for

both operation modes. Hence, the amount of propane entering the C3 product stream can

be calculated from the amount ofthe propylene entering the C3 product stream using the

formula given below:

V - ^PC TT/ _ propane ' propane ^ g 2^)

propane y _ ^ n ^ y ^p^ ' \ • ) propane propane propylene propylene

W = I-W (6 3\ propylene propane'' v"--"/

where

Wpropane, Wpropyiene- weight percentages of propane, propylene in the C3 product,

respectively,

Vpropane, Vpropyiene- volumc percentages of propane, propylene in the C3 product,

respectively.

The densities of propane and propylene are from literature (Edmister and Lee,

1984).

The mass flow rate ofthe propylene can be calculated from its weight percentage

and the mass rate ofthe propane using the formula given below:

W WT =WT ._p^£py!f!!L (f.A) ' propylene propane rry ' v " ' ^ ^

propane

where

WTpropane, WTpropyiene- mass flow ratcs of the propanc and propylene in the C3 product,

respectively.

The mass flow rates of propane and propylene in the alkylation feed are obtained

by subtracting the propane and propylene in the C3 product from the total propane and

propylene in the feed to the depropanizer.

125

Page 141: 5-95 gap chart

6.2 Alkylation Unit

In the alkylation unit, the propylene and butylene react with isobutane to form

isoheptane and isooctane, respectively. Isoheptane and isooctane are two major

components ofthe alkylate. Isooctane and isoheptane are excellent gasoline blending

stocks with high octane number and low Reid Vapor Pressure (RVP). The normal

paraffins, such as propane and normal butane, do not react in the alkylation unit. They are

separated from the alkylate in the separation train ofthe alkylation unit.

Although the alkylation reactions can take place at high temperature and pressures

without catalysts, the only processes of commercial importance involve low-temperature

alkylation conducted in the presence of either sulfuric or hydrofluoric acid (Gary and

Handwerk, 1984). The refinery considered in this work uses hydrofluoric acid as the

catalyst in the alkylation unit.

In the alkylation unit, high isoparaffin/olefin ratios (4:1 to 15:1) are used to

minimize polymerization and to increase product octane (Gary and Handwerk, 1984).

The isobutane produced inside the refinery is not enough to satisfy the requirement ofthe

alkylation reactions. Isobutane purchased in the market is added to alkylation feed before

it enters the alkylation unit.

A simplified schematic ofthe alkylation unit is shown in Figure 6.2. In the

reaction part ofthe alkylation unit, the alkylation feed is mixed with hydrofluoric acid in

an acid settler. Since the acid phase is heavier than the hydrocarbon phase, the reaction

mixture settles into two liquid layers in the settler. The acid is withdrawn from the bottom

ofthe settler and then recycled and mixed with more fresh feed. The hydrocarbon layer is

a mixture of propane, isobutane, normal butane, and alkylate. The hydrocarbon layer is

withdrawn from the top ofthe settler and flows to a depropanizer.

126

Page 142: 5-95 gap chart

O 3

0)

3 K . 5 .N a> C

X)

Q

td p t-(u F y ^ o «

OH I>

0) c CO a. o

C N

i> S

c >~ 5

•O O 'o U

<

o x> <u o

<u f )

>.

Rec

c m ^

sob

^ 1/3 • f

<u H-» 3

XI O

c .2 •5 "^

CO ai

^ fa

<

X3 T3

£ <

c

c o

'C o

o

ca

o

ca a (U

J3 o

1)

S)

127

Page 143: 5-95 gap chart

The propane is withdrawn from the top ofthe depropanizer and becomes a final

product. The bottom stream withdrawn from the depropanizer goes to a deisobutanizer.

The isobutane exits the deisobutanizer from the top and is recycled to react with more

fresh feed. Normal butane and alkylate exits the deisobutanizer from the bottom and

flows to debutanizer. The butane is withdrawn from the top ofthe debutanizer and goes

to the gasoline blending unit as a blending stock. The alkylate is withdrawn from the

bottom ofthe debutanizer and goes to the gasoline blending unit as a blending stock.

A simplified model was developed for the alkylation unit. The model is based on

the stoichiometric relations ofthe alkylation reactions. Three major alkylation reactions

are given below:

CH3 CH3 CH3 CH3

CH3-C=CH2 + CH3-CH-CH3 -^ CH3-C-CH2-CH-CH3

CH3

Isobutylene Isobutane 2,2,4-trimethylpentane (Isooctane)

CH3 CH3 CH3

CH3-CH2-C=CH2 + CH3-CH-CH3 -» CH3-CH2-C-CH2-CH-CH3

butylene Isobutane 2,4-dimethylhexane

: H 3 CH3

CH3=C-CH3 + CH3-CH-CH3 ^ CH3-C-CH2-CH2-CH3

Propylene Isobutane 2,2-dimethylpentane (Isoheptane)

The first two reactions can be considered as one reaction in the model since the

properties ofthe reactants and the product of both reactions are very similar since they

are isomers. The model does not distinguish between the normal butylene and the

isobutylene in the alkylation feed.

128

Page 144: 5-95 gap chart

Another path in propylene alkylation is the combination of propylene with

isobutane to form propane plus isobutylene. The isobutylene then reacts with more

isobutane to form isooctane (Gary and Handwerk, 1984). The reaction of propylene with

isobutane is shown below:

CH3 CH3

CH3=C-CH3 + CH3-CH-CH3 -^ CH3-CH2-CH3 + CH3-C=CH2

Propylene Isobutane Propane Isobutylene

It is assumed that all the propylene and the butylene in the alkylation feed react

with isobutane to form the alkylate. Since there is only one reaction path for the butylene,

the amount of isobutane required and the amount of isooctane produced from butylene

can be calculated from the stoichoimetric relation ofthe butylene alkylation.

There are two reaction paths for the propylene to form alkylate. The amount of

propylene for each reaction path needs to be determined. The weight percentage of

propylene that reacts in the second reaction path based on industrial data is listed in the

Table 6.5. The difference between two modes is 0.0029 wt.%. hence, the average value

listed in Table 6.5 is used for both operation modes in the model. It is assumed that this

average weight percentage is constant and does not change with operating conditions.

Table 6.5 Weight Percentage ofthe Propylene that Reacts in the Second Reaction Path in the Alkylation Unit.

Propylene going though the second reaction path, wt.%o

Summer Mode 0.0458

Winter Mode 0.0429

Difference 0.0029

Average 0.0444

The amount ofthe isooctane and the isoheptane in the alkylate can be calculated

from the stoichoimetric relations. Since all the stoichoimetric constants in all the

129

Page 145: 5-95 gap chart

alkylation reactions are equal to unity, the stoichoimetric constants do not appear in the

equations of calculating the isooctane and the isoheptane in the alkylate. The calculations

use the following formula:

W • MW W -X . .-MW j-fT _ "butylene ^'-"'isooctane , "propylene ^'^ sec ond path •* i.sooctane (f,A\

isooctane ~ H/flJ/ A/fW ' V • / ^ ' ^ butylene ^ ' ^ propylene

„•• propylene \ second path) ^^ isoheptane rr r\ isoheplane i rwrr ' v * /

MW I propylene

where

Wisooctane, Wisoheptane" Hiass flow ratcs of isooctauc and isoheptane, respectively, MLB/D,

Wbutyiene, Wpropyiene- mass flow ratcs of butylcnc and propylene in the alkylation feed,

respectively, MLB/D,

MWisooctane, MWisoheptane, MWbutylene, MWpropylene" m o l c C u l a r WCightS o f isOOCtaUC,

isoheptane, butylene, and propylene, respectively, g/mole,

Xsecond path- Weight fraction of propylene that reacts in the second reaction path.

The alkylate formed by butylene is called butylene alkylate and the alkylate

formed by propylene is called propylene alkylate. The octane number ofthe butylene

alkylate and the propylene alkylate are listed in Table 6.6. These numbers come from

Gary and Handwerk (1984).

Table 6.6 Octane Number ofthe Butylene AH

Type RON

Butylene alkylate 96

Propylene alkylate 93

kylate and Propylene Alkylate.

MON

94

91

The octane number ofthe alkylate is calculated as the volumetric average value

using the formula given below:

130

Page 146: 5-95 gap chart

alkylate butylene alkylate propylene alkylate

propylene alkylate propylene alkylate ' V • /

where

Ouaikyiate- MON Or RON ofthe alkylate product,

Oubutyiene alkylate, Oupropyiene alkylate" MON Or RON of the butylcne alkylate and the

propylene alkylate, respectively,

Vbutyiene alkylate, Vpropyiene alkylate- volumctric fractions of the butylcnc alkylate and the

propylene alkylate in the alkylate product stream, respectively.

The volumetric fractions ofthe butylene alkylate and the propylene alkylate are

calculated from corresponding mass flow rates and the densities ofthe two components.

The amount ofthe isobutane required in the reactions can be calculated using the

formula given below:

W - HfW J-.-, _ butylene isobutane

isobutane, reaction A/fJJ/ ^ ' ^ butylene

propylene ( J ' - ^ second path / - ' - ' ' ' isobutane . , _ .

propylene

where

Wisobutane, reaction- mass flow rate of the isobutane required in the alkylation reactions,

MLB/D,

MWisobutane- molccular weight of isobutane, g/mole.

It should be noted that one mole of propylene that reacts in the second path

consumes two moles of isobutane. Hence, one mole of propylene in the alkylation feed

consumes (1+Xsecondpath) moles of isobutane.

131

Page 147: 5-95 gap chart

The isobutane can not be completely separated from the normal butane in the

deisobutanizer. Some isobutane enters the mixed butane product stream. The weight

percentage of the unreacted isobutane in the total isobutane in both operation modes is

obtained from the industrial data from the refinery considered in the work. These values

are listed in Table 6.7. It can been seen from Table 6.7 that the difference between two

modes is only 0.0003 wft.%. Hence, the average value is used for both operation modes in

the model and it is assumed that the weight percentage does not change with the

operating conditions ofthe alkylation unit.

Table 6.7 Weight Percentage ofthe Unreacted Isobutane in the Alkylation Unit.

Unreacted isobutane, wt.%)

Summer Mode

Winter Mode

Difference

Average

0.0211

0.0208

0.0003

0.0209

The total amount ofthe isobutane required by the alkylation unit is calculated

using the formula given below:

W - J . isobutane. reaction ,-. Q , " isobutane. total ~ ~, 7^ ' t ^ - O j

isobutane, unreacted

where

Wisobutane, total- total mass flow rate of isobutane required by the alkylation unit, MLB/D,

Xisobutane,unreacted- weight fraction of the unrcacted isobutane in the total isobutane required

by the alkylation unit.

The volumetric flow rate ofthe isobutane is calculated from the mass flow rate

and the density of isobutane. The amount ofthe isobutane purchased on the market is the

132

Page 148: 5-95 gap chart

calculated as the difference between the total isobutane required by the alkylation unit

and the available isobutane in the alkylation feed:

W =W -W (6.9) isobutane. purchased isobutane. total isobutane. feed ^ ^ ^

where

Wisobutane, purchased- mass flow rate of isobutauc purchased in the market, MLB/D,

Wisobutane, feed- mass flow rate of isobutanc in the alkylation feed, MLB/D.

The total amount ofthe mixed butane is the sum ofthe butane in the alkylation

feed and the unreacted isobutane calculated using the following formula:

W =W -^W. - X ('6 10) bulane.lolal normal butane isobutane .total isobutane .unreacted ' V • /

where

Wbutane, total" mass flow rate of mixed butane, MLB/D.

The volumetric flow rate ofthe mixed butane is calculated from the mass flow rate and

the density of butane.

The amount ofthe propane formed in the second path of propylene can be

calculated using the formula given below:

W - X -MW „ . propylene second path propane /"/: 1 1 \

propane. reaction h/fJJf/ ' V ^ * ^ ^ / IVI rr p^QpyigfjQ

where

Wpropane, reaction- mass flow rate of propauc formed in the alkylation unit, MLB/D,

M Wpropane- molccular weight of propane, g/mole.

133

Page 149: 5-95 gap chart

The total mass flow rate ofthe propane is the sum ofthe propane in the alkylation feed

and the propane formed in the alkylation unit. The volumetric flow rate ofthe propane is

calculated from the mass flow rate and the density of propane.

6.3 Diesel Hydrotreater

The purpose ofthe diesel hydrotreater is to eliminate most ofthe sulfiir in the

diesel to satisfy the diesel product specification. The diesel produced in the refinery

considered in this work is composed of three diesel streams. These streams are the side-

draw diesel from the atmospheric tower in the cmde unit, the overhead diesel from the

remn colunm in the cmde unit, the light cycle oil (LCO) from the main fractionator of the

FCC unit. The mass flow rates and volumetric flow rates of these three streams have been

calculated in the cmde unit model and FCC model, respectively. The diesel hydrotreater

model calculates the hydrogen required and the quantities ofthe product streams.

There are three product streams coming out ofthe diesel hydrotreater: diesel,

naphtha, and sour gas. The diesel goes to the storage tank. The naphtha goes to the cmde

unit and is combined with the light naphtha stream. The sour gas goes to Amine treater

for further treatment. The average values ofthe industrial data from the refinery

considered in this work shown in Table 6.8 and Table 6.9 are used in the model. It can

been seen from Table 6.8 and Table 6.9 that the values for the summer mode and the

winter mode are close, with the largest difference of 0.003 in weight ratios and 0.005 in

volumetric ratios. Hence, the average values are used for both operation modes in the

model. These ratios are constants and do not change with the operating conditions.

Table 6.8 Weight Ratios ofthe Hydrogen and the Products in the Diesel Hydrotreater.

Summer Mode Winter Mode Difference Average

Hydrogen/Charge 0.018 ~"

Diesel/Charge 0.970

Naphtha/Charge 0.024

Sour Gas/Charge 0.025

0.018

0.973

0.022

0.022

0.0

0.003

0.002

0.003

0.018

0.971

0.023

0.024

134

Page 150: 5-95 gap chart

0.026

0.983

0.029

0.015

0.001

0.005

0.001

0.003

0.027

0.981

0.030

0.016

Table 6.9 Volumetric Ratios ofthe Hydrogen and the Products in the Diesel Hydrofreater.

Summer Mode Winter Mode Difference Average

Hydrogen/Charge 0.027 ~

Diesel/Charge 0.978

Naphtha/Charge 0.030

Sour Gas/Charge 0.018

Since the diesel hydrotreater has a capacity of 13,000 barrels charge per day, not

all diesel streams go through the diesel hydrotreater. Part of diesel withdrawn from the

atmospheric tower in the cmde unit directly goes to the storage tanks without the

treatment. About 20 vol.% in the Winter Mode and 24 vol.% in the Summer Mode of

diesel from the atmospheric tower do not go through the treatment in the diesel

hydrotreater. These values are used as constants in the model directly.

The refinery considered in this work produces two diesel products: low sulfur

diesel fuel and No. 2 diesel fiiel. There is not a strict limit on the quantities of no. 2 diesel

and low sulfur diesel. In the winter, the market demand for low sulfur diesel is high.

Thus, maximum low sulftir diesel is produced while reaching the upper limit of diesel

hydrotreater capacity. The rest ofthe diesel which is not processed by the diesel

hydrotreater is sold as No. 2 diesel fuel. In the summer, however, the low sulfiir diesel

demand is down and there is no need to reach the HDS limit to produce the maximum

amount ofthe low sulfiir diesel. The low sulfur diesel is about 76 vol.%) ofthe entire

diesel produced in the refinery. The rest ofthe diesel, whether processed by the diesel

hydrotreater or not, is sold as No. 2 diesel fiiel. That portion of low sulfur diesel in the

summer, 76 vol. %, is assumed as a constant in the model.

135

Page 151: 5-95 gap chart

CHAPTER 7

GASOLINE BLENDING MODELING

7.1 Process Overview

Gasoline is the primary product in a fuel-oriented refinery. In the refinery

considered in this work, the volume ofthe gasoline is about half of the total volume of all

products produced in the refinery. From the point of view of economics, about 60-70%) of

a typical refinery's total revenue comes from the gasoline sale (Singh et al., 2000).

Gasoline blending is the final step in making gasoline product. The gasoline blending

operation often determines the operating conditions ofthe upperstream units. Due to the

importance ofthe gasoline blending, a gasoline blending model must be included in the

refinery-wide model.

Gasoline blending is the process of blending several gasoline blending stocks that

are produced in upperstream units or purchased from the market to make several grades

of gasoline according to the specifications. The objective ofthe gasoline blending is to

allocate the available gasoline blending components in such a way as to meet product

demands and specifications at the least cost and to produce products which maximize the

overall profit. Different gasoline blending stocks have different properties. Different

grades of gasoline also have different specifications. The core of a gasoline blending

model is the prediction of gasoline properties from the properties ofthe blending stocks.

Table 7.1 lists the volumes and sources ofthe gasoline blending stocks in the

typical Summer Mode operation. In Table 7.1, the first seven blending stocks are

produced and blended in the refinery while the ethanol is off-site. The high solvability of

the water in ethanol makes it inconvenient to blend ethanol in the refinery and transport

the gasoline with ethanol in it. Water may enter the gasoline during the transportation.

It can be seen from Table 7.1 that the refinery itself makes more than 97 vol.%

percent of gasoline blending stocks. Therefore, adjusting the operating conditions of

upstream units according to the gasoline blending is essential to make the refinery

operation profitable.

136

Page 152: 5-95 gap chart

Table 7.1 Volumes and Sources ofthe Gasoline Blending Stocks in Typical Summer Mode Operation.

Gasoline blending stock Volume, BBL/D Vol. Vo Source

FCC gasoline 10915

Low-severity reformate 5934

Alkylate 4000

Light straight-run gasoline 1783

High-severity reformate 1244

Butane 288

Toluene

Ethanol

200

379

44.11349

23.98254

16.16619

7.206078

5.027685

1.163966

0.808309

1.531746

FCC unit

Reformer

Alkylation unit

Cmde unit

Reformer

Gas plant and/or

market purchase

Market purchase

Market purchase

The refinery makes three grades of regular gasoline: super unleaded (SNL),

unleaded (NL), and sub. octane with ethanol. The super unleaded gasoline has an octane

rating of 93 and the unleaded gasoline has an octane rating of 87. The sub. octane with

ethanol gasoline has an octane rating of 84 inside the refinery and its octane rating is

increased to 87 when ethanol is blended into the gasoline at the pump. The refinery does

not make any reformulated gasoline (RFG) at present. The specifications of these three

grades of gasoline are listed in Table 7.2.

The octane number of a fuel is defined as the percentage of iso-octane (assigned

an octane number of 100) in a blend with n-heptane (assigned an octane number of 0) that

exhibits the same resistance to knocking as the test fuel under standard conditions in a

standard engine (Palmer and Smith, 1985). Two standard test procedures are used to

characterize the antiknock properties of fuels for spark engine: (I) the ASTM D908 test

gives the research octane number (RON) and (2) the ASTM D357 gives the motor octane

number (MON). The RON represents the antiknock properties under the condition of low

137

Page 153: 5-95 gap chart

ork.

^ c/n

15 H . f i

1)

der

tzi fi O

u

nery

i n

n th

e R

e uc

ed i

- o o V-i

olin

e

CO

•fga

c/3

;e G

rad

.5 H Ct-I o fi o

ta o

«

eci

o , in

ble

7.2

ca H

I-I

H-»

c

I - I

c 3

<U fi ca

H-» CJ

o

Sub.

ad

ed

nle

:D

T3

lea(

fi ^

I - I

<u OH 3

C/3

Oct

ane

Sub.

ad

ed

u

Unl

le

aded

fi

I - I

D OH

fi _o

ecif

icat

OH

pper

P

Low

er

Upp

er

I—.

owei

H J

pper

P I - I

Low

l U

pper

ow

er

hJ

Upp

er

Low

er

pper

P ^ (L>

o H-l

"a

limit

limit

H-» . .—1

a *

1

.H-»

lim

limit

mit

• i-H

limit

-t-»

lim

1

s—»

1

fi o :z;

CN 00

c o

7.

CN 00

fi o 7.

m fi o 7.

Non

e

CN 00

fi O

"7,

(N 00

Non

e

u fi o 2

MO

N

fi o

2

t ^ oo

fi o

7

00

fi o

2

r ' l ON

Non

e

t ^ 00

fi O

"7,

oo

Non

e

m as

)/2

2 O

ON

+M

6

»n r<i

6.4

U-) CN

6.4

m CN

^ vd

oo od

6.4

oo t~-

6.4

7.8

'^ so

p, p

sia

^

00

0.7

0.72

00

0.7

CN t ^ O

00

0.7

CN t ^

d>

0.78

4 0.

72

oo

0.7

CN t ^ O

0.78

4

CN t ^ O

H-»

> ca I-I

ecif

ic g

OH

1200

N

one

o o CN

u fi o 7

o o CN

u fi o 7

1250

N

one

o i n CN

one

7

1250

<u fi o 7

X

T3 fi

•^ H-»

ivea

bili

I-I

Q

^ m

Non

e

^ m

(U fi o

Z

,_, m

u fi o 2

158

Non

e

oo i n

one

7:

oo m

u fi o

:^

•*-J

w

Q ^ o^ O ^

•lat

ility

.

>

i n m CN

170

i n

O

'"—'

i n m CN

o t ^

250

170

o i n (N

o

250

o t ^ '"'

H-* C«

b O i n

ilatil

ity,

>

>n

rn

Non

e

i n so m

<a c o

7

i n

U fi O

:^

374

Non

e

^

one

2

374

OJ fi o 2 • 4 - * C/3

• i - H

Q ^ o^ O as

ilatil

ity.

>

O ^

3.7

r^ (N

U C o

z t ^ CN

u fi o

z

4.0

3.7

t~-CN

one

7

2.7

u c o 2

: wt.%

CN

o

^.^

cd

Non

e

o

u c o

2

o

u c o 7.

0.1

Non

e ^ o

one

7

O.I

u C o 2

^

Iftir

, w

t

3 or)

ON 'S-'

Non

e

as ^

<\i fi o

7

as ^

u fi o 7,

4.9

Non

e

as •^

one

2

4.9

u fi o 2

^

i

nzen

e.

CQ

138

Page 154: 5-95 gap chart

speed and frequent accelerations while the MON represents the engine performance

under more severe high speed conditions (Singh et al., 2000). In RON test, the engine

runs at 600 r/min and with 125 °F intake air temperatiire. In MON test, the engine mns at

900 r/min and with 300 °F intake air temperatiire (McKetta, 1992). The arithmetic mean

ofthe RON and MON, (R0N+M0N)/2, is called octane rating and is posted at gasoline

stations. Both the octane rating and MON are used in the gasoline specification.

Automobile engine performance is affected by gasoline volatility. A vapor

pressure that is too high for the given ambient temperature will result in vapor locking

and motor stalling, while a vapor pressure which is too low will lead to difficulties in

engine start-up (Palmer and Smith, 1985). Reid vapor pressure (RVP) is widely used as a

criterion to measure the volatility of gasoline. RVP test is defined by the American

Society for Testing and Materials (ASTM) under the designation ASTM D323-56. The

American Petroleum Institute (API) describes the RVP test procedure in detail, including

the apparatus (API, 1955).

The boiling range also affects the engine during start-up and driving, and is

particularly important for good performance during quick acceleration and high speed

operations (Singh et al., 2000). The 10, 50, and 90 vol.% distilled are used in the gasoline

specifications. The 10 vol.% distilled affects the start-up and vapor blocking. The 50

vol.%) distilled affects the acceleration and smoothness. The 90 vol.% represents the

completeness of gasoline combustion (Lin, 1988). ASTM D86 standard test is used to

measure the 10, 50, and 90 vol.% distilled ofthe gasoline.

The volatility of fuels is varied for seasonal climatic changes and conformance to

U.S. EPA volatility regulations by providing six vapor pressure/distillation classes and

six vapor lock protection classes for fuel (ASTM, 1997). The volatility of fuels is also

varied for different geographical regions. According to the market locations ofthe

refinery considered in this work, the volatility specifications of gasoline grades SI and S4

are applied for RVP and 10, 50, and 90 vol.% in the summer and the winter, respectively.

It is to be noted that the upper limit of RVP ofthe gasoline sub. octane with ethanol is I

psi higher than other two grades of gasoline. The additional 1 psi RVP is an incentive to

139

Page 155: 5-95 gap chart

use the replaceable ethanol in the gasoline. The ethanol content in the gasoline sub.

octane with ethanol is about 10. vol.%.

The auto and oil industry accepts a driveability index (DI) based on the ASTM

D86 distillation curve to protect against the performance problems related to volatility. It

is suggested that the DI should not exceed 1,200 to assure satisfactory performance

(Unzelman, 1996). The Driveability index is calculated using the following formula:

DI = I.5-Tj„+3-T,„+T,o, (7.1)

where

DI- driveability index,

Tio, T50, T90- temperature at 10, 50, and 90 vol.%) distilled, respectively, °F.

The specification ofthe O2 wt.% is to control the oxygen content in the gasoline.

The super unleaded gasoline and the unleaded gasoline are conventional gasoline. The

gasoline sub. octane with ethanol is an oxygenated gasoline. The lower limit of oxygen

wt. % for the oxygenated gasoline is defined in 1990 Amendment to the clear air act. The

upper limit of O2 wt.% is set to control the Nox emission from the gasoline. It is noted that

the upper limit of O2 wt.% for the super unleaded gasoline and the unleaded gasoline is

lower than that of gasoline sub. octane with ethanol.

The specific gravity specification is to control the density ofthe gasoline. The

lower limit of specific gravity can prevent the gasoline from getting too light. A heavy

gasoline usually yields better mileage than a lighter gasoline. In gasoline blending, the

lower limit is mainly used to limit the content of alkylate, a light high-octane gasoline

blending stock, in the super unleaded gasoline. However, a gasoline that is too heavy

usually has low volatility, which may cause ignition problem. A upper limit of specific

gravity is also set in the gasoline specifications.

The sulfur specification is to limit the emission of SOx from the gasoline. SOx is

harmful to the environment and considered to be responsible for acid rain. The upper

limit of benzene wt. %> in the gasoline specifications is to limit the emission of CO from

140

Page 156: 5-95 gap chart

the gasoline. Since the hydrogen/carbon ratio is low in benzene, gasoline with high

content of benzene tends to have high emission of CO.

The benzene content, O2 content, and sulfur content can be easily calculated by

summing the corresponding contents in the gasoline blending stocks. The specific gravity

ofthe gasoline can be easily calculated from the volume and the weight ofthe gasoline. It

is assumed that there is no volume loss in gasoline blending. Hence, the volume of one

grade of gasoline is the summation ofthe volumes of gasoline blending stocks blended

into the gasoline. The weight ofthe gasoline is calculated by summing the weights of

gasoline blending stocks blended into the gasoline.

The calculations of octane number, RVP, and percent distilled are more

complicated and will be described in detail in the following text. After the 10, 50, and

90%) distilled are calculated, the driveablity index can be calculated using equation 7.1.

7.2 Properties ofthe Gasoline Blending Stocks

In order to calculate the properties ofthe gasoline blends, the properties ofthe

gasoline blending stocks must be known. Some properties ofthe gasoline blending stocks

are calculated in the unit models. Average industrial data are used for other properties.

The properties ofthe blending stocks required by the gasoline blending model includes

twelve items: volumetric flow rate, mass flow rate, MON, RON, saturates

(paraffin+naphthene) vol.%), aromatics vol.%, olefins vol.%), RVP, ASTM D86 curve,

benzene w^.%, sulfur wt.%, and O2 wt.%). There are all together eight blending stocks:

light straight-run (LSR) gasoline, FCC gasoline, low-severity reformate, high-severity

reformate, alkylate, butane, toluene, and ethanol. The eight blending stocks will be

introduced in the following text.

7.2.1 Light Straight-Run (LSR) Gasoline

Recall that in the cmde unit the overhead product of naphtha splitter becomes the

LSR. The LSR flows direct from the crude unit to the gasoline storage tank for blending

without going through any treatment in between. Therefore, the properties ofthe LSR can

be calculated directly from the information ofthe portion ofthe cmde feed in the LSR

141

Page 157: 5-95 gap chart

boiling range, which are listed in Table 7.3 and Table 7.4. The LSR consists the whole

gasoline cut and part of light naphtha cut in the cmde feed. The rest of light naphtha cut

becomes reformer feed. The portion of light naphtha cut entering the LSR is calculated

using the formula given below:

Portion,,^,„ = 1 - Portion ^j„^,^^j^^,,„, (7.2)

where

PortiouLSR, In- volumetric fraction ofthe light naphtha in the LSR,

Portionreformer feed. In" volumctric fraction of the light naphtha in the reformer feed.

Recall that the Portionreformer feed, In is calculated in the reformer model.

Table 7.3 Industrial Data of Cmde A and Cmde B.

Property Unit Cmde A CmdeB

Item Gasoline

C5-150°F

Light naphtha 150-200°F

Gasoline

C5-150°F

Light naphtha I50-200°F

Yield

API Gravity

Aromatics

Naphthenes+2x

Aromatics

RON

MON

Benzene

Sulfiir

Vol.%

API

vol.%

vol.%

None

None

wt.%

wt.%

2.5

88.687

0.223

6.447

72.479

71.684

O.I

0.001

2.5

67.69

1.515

37.93

71.832

67.992

4.2

0.007

3.1

91.041

1.151

5.705

70.325

69.318

O.I

0.003

1.9

73.961

8.556

40.037

71.832

67.992

4.2

0.013

142

Page 158: 5-95 gap chart

Table 7.4 Indusfrial Data of Cmde C and Cmde D.

Property

Item

Yield

API Gravity

Paraffins

Naphthenes

Aromatics

RON

MON

Aromatics

Sulfiir

Unit

Vol.%

API

vol.%

vol.%

vol.%

None

wt.%

wt.%

Wt. %

CmdeC

Gasoline

68-155 °F

3.98

86.5

92.8

6.2

0.1

69.5

68.3

1.0

<.02

Light

naphtha

155-265 °F

10.33

63.1

62.3

31.9

4.2

-

-

5.8

<.02

CmdeD

Gasoline

C5-150°F

2.88

83.6

90.8

6.6

O.I

71.4

69.8

2.7

0.02

Light

naphtha

150-200 °F

9.18

60.9

57.4

34.3

4.2

-

-

8.3

0.02

The volumetric flow rate and the mass flow rate ofthe LSR are calculated in the

naphtha splitter part ofthe cmde unit model. The MON and RON ofthe LSR can be

calculated using the formula given below:

2^/oVi - \ygasoline,i ' ^gasoline,! "^ "^^^^^^LSR.In.i ' '^ln,i ' ^ln,i )

Y _i ~ E ^ ^ / • (^gasoline,i + Portion^^j,,,,, - V,„, ) '

(7.3)

where

i- cmde types of A, B, C, and D,

Y- MON or RON,

%Vi- volume percentage of crude type i in the total cmde, vol. %,

Vgasoiine,i- volumc fraction ofthe gasoline cut in the cmde type i,

143

Page 159: 5-95 gap chart

Vin,i- volume fraction ofthe light naphtha in the crude type i,

Ygasoiine,i- MON Or RON of the gasoline cut in the cmde type i,

Y|n,i- MON or RON ofthe light naphtha in the cmde type i.

It can be seen from Table 7.4 that the MON and RON ofthe light naphtha cut in

the cmde C and cmde D are not available in the industrial data. It was assumed that the

MON and RON ofthe light naphtha cut in these two cmde types are the same as the

MON and RON ofthe gasoline cut in the same cmde type, respectively.

The saturates include paraffins and naphthenes. The volume percentages ofthe

saturates in the gasoline cut and light naphtha cut ofthe cmde A and cmde B can be

calculated by summing corresponding volume percentages ofthe paraffins and

naphthenes. The volumetric percentage ofthe olefins in the gasoline cut and the light

naphtha cut are set to zero based on industrial data. The volume percentages ofthe

saturates, the aromatics, and the olefins in the LSR can be calculated from the volume

percentage of these components in the gasoline cut and light naphtha cut of four cmde

oils using the formula given below:

X ^ ^ ' • (^gasoline.i " ^x.gasoline.i + Portion,^„j„j - V,„j - V^j„)

"^oX = ^ v^TTTT-rr ^T-TT 7^ ' (7-4) Y^oV. • (V^asoiinci + Portion,SR,,„„ • V,J

where

i- cmde type A, B, C, and D,

X - component type, saturates, aromatics and olefins,

%X- volume percentage of component type X, saturates, aromatics and olefins, vol.%,

Vx,gasoiine,i- volumc fraction ofthe molecule type X in the gasoline cut ofthe cmde type i,

Vx,in,i- volume fraction ofthe molecule type X in the light naphtha ofthe cmde type i.

The benzene content and the sulfur content in the LSR can also be calculated from

the corresponding contents of gasoline cut and light naphtha cut of four cmde oils using

the formula given below:

144

Page 160: 5-95 gap chart

Y:/ow, *^gasoline.! ''^"^gasoline.i '" Z.gasoline.i ''^ "0rtl0n[^n,^j

V^SPG,„rW.. %Z = - ^ - ^ '^—^-^ -_ ^ '-, (7.5)

^Vgasoline.i ' ^PGga^oline.i + PortiOn,^^,„, "^

KVln.i-SPG,„^,

where

i- cmde types of A, B, C, and D,

Z - benzene or sulfur,

%)Wi- weight percentage of cmde type i in the total crude, wt. %,

%Z- benzene content or sulfur content, wt.%,

Wz,gasoiine,i- weight fraction ofthe benzene or sulfiir in the gasoline cut ofthe cmde type

i,

Wz,in,i- volume fraction ofthe benzene or sulfiir in the light naphtha ofthe cmde type i.

The RVP, and the O2 wt.% in the LSR are assumed to be constants in the model.

They are set at the average industrial data. According to the operation personnel in the

refinery considered in this work, the ASTM D86 curves of gasoline blending stocks

change little during the operation. Hence, the average industrial values are also used for

the ASTM D86 curve ofthe light naphtha. These constants are listed in Table 7.5.

7.2.2 Other Gasoline Blending Stocks

The mass flow rate, MON, and RON ofthe FCC gasoline are calculated in the

FCC model introduced in the Chapter 4 of this dissertation. However, the FCC model

does not predict the volumetric flow rate ofthe FCC gasoline. The API gravity ofthe

FCC gasoline is assumed to be a constant and is set at the average API gravity ofthe FCC

gasoline in the industrial data. The FCC model can not predict the volumetric percentages

ofthe chemical components in the FCC gasoline, either. The industrial average values are

used for the volumetric percentages ofthe saturates, the aromatics, and the olefins in the

FCC gasoline. The RVP, ASTM D86 curve, benzene wt.%, the sulftir wt.%, and the O2

145

Page 161: 5-95 gap chart

wt.% in the FCC gasoline also can not be predicted from the FCC model. Their values are

set at the average industrial data.

For both the low-severity operation and high-severity operation ofthe reformer,

the detailed reformer model provides the volumetric flow rate and the mass flow rate of

the reformate. The reformer model also provides the contents ofthe saturates, the

aromatics, and the olefins in the reformate. In addition, the model provides the benzene

content in the reformate. The RVP, ASTM D86 curve, sulfur wt.%), and the O2 wt.% in

the reformate can not be predicted from the refomate model. Their values are set at the

average industrial values.

The alkylation model provides the volumetric flow rate, mass flow rate, MON and

RON ofthe alkylate. Other properties are not calculated in the model and they are set as

the average industrial values.

The volumetric flow rate and the mass flow rate ofthe butane are calculated in the

gas plant model. Other properties ofthe butane are obtained from literature (Edmister and

Lee, 1984; Gary and Handwerk, 1984) and from industrial data.

While all the gasoline blending stocks introduced above are produced inside the

refinery, the toluene and the ethanol are only purchased from the market. The refinery

also purchases some butane from the market in the winter. The butane purchased is

assumed to have the same properties ofthe butane produced in the refinery. The

volumetric flow rate and mass flow rate ofthe toluene and the ethanol are assumed to be

at the upper limit ofthe purchase at the beginning ofthe calculations. The real amount of

the toluene and the ethanol used are calculated in the gasoline blending model. Other

properties ofthe toluene and ethanol are obtained from literature (Edmister and Lee,

1984; Unzelman, 1996) and from industrial data.

146

Page 162: 5-95 gap chart

CO

M o o

H-»

c

fi

s _fi

CO

ca O o CO

<L)

u OH O

i n t~~ u

u C <L>

_3 "o

(U fi

3 CQ

j=l 00

CO K

h-1

i n CN CN i n ^ vd >^ OO CN ON 00 T t ^ ON O O O —I

t ^ o i n ^ O r n o ^ o CN o o n m - ^ ^ o T^ o o

ON

d ^ - ^ o >—I r<i o o r s , -—1 ON ON ^ ^ O O i n

o ?: <u u C^

S S S as o od C3N

m ^ vo • ^ t ^ ON 0 0 c ^

^ g : 9 t< ^ o d ^ . " : ^ . vo m i n ON —I

O O vo ON ^ —I ^

s s s s s s

i n CN i > i n S2

vq ' ^ CN m r~

t; OH O I-I

( IH

2; o

"o >

ta

•^ Pii S c/3 ^ o

"o >

fi

u

ca ^ • ^ o

PH H > 03

O

H 00 m 00

O i n

H 00 <

o as i n

ON

H 00

<

CN

t ^ — o

9 9 - ^ o m

m m o o I CN - ^ O O

^ p p p m o o d

O N - ^ v o c N O o o m o o N m i n t ^ - r t o m m c N v o v o

S t ^ c J N ^ ^ ' c N ' ^ i ^ v d - ^ ' ^ ' ^ ' ^ ' ^ ^ ' ^ • 0 ^ t ^ ( ^ t ~ - 0 ' * t ^ C N

- ^ m r O C N V O O N V O V D C N ^ ' ~ ~ : ' ~ i ^ ^ ^ ' ^ ° ® ^ f ^ L 2 — ' i n — i v d i n d r < S r n i n

' ^ ' ^ ' ^ ' ^ ' ^ ' ^ . — ' c o ' ^ i n t ^ o - ^ t ^ o l ^ i ^ P « S l ^ i « ! 5 ( < t ^ ' - H r - l , — I . — I , — l ( N | ( N l C N m I

o o o d d

ss

O t ^ ON 00 o ON vq i n ON ON

t~~ t~~ ON t~-^ CN t ^ m ON

CN CN CN r n m

v O ' ^ v o t ^ v o o N C j N O N i n t ^ v q p c N i n r n o ^ ^ O - * S ^ ' c N r < - i v d v d c N v d v d t ~ ~ ^

. o i n t - ~ c N v o o - * t ^ ^ \ 0 , - H ^ H ^ H ( v ) r N l r < ^ m r < - ) H ^

• ^ o o t ^ O N i n ' ^ t ^ o T t ' v t ^ ^ m o o ^ H ^ H O v i n ^ ^ ^ " ^ r n ' ^ ' d ' S ^ d o d — J c N

^ CN ON vo o CN

o p o o CN d d

o o o d d

00 oo o o d) d> <d'

tin O

ON. O o _ ^ CL, 5 00 r^ fi I i | s

O 1) CQ CQ 00 O

ID -o o

a -o

-a <L> I-I O H

CO

_3 ta > CO

- f i .22

"o,

H-t o c

_3 l a > CO

.a ca

J3

<u

ca J3

I i — CN

147

Page 163: 5-95 gap chart

All these industrial data and literature data for the properties ofthe gasoline

blending stocks mentioned above are listed in Table 7.5. After all the properties ofthe

gasoline blending stocks are known, they are used as inputs to the gasoline blending

model.

7.3 Gasoline Blending Model

7.3.1 Octane Model

The Octane Model is to calculate the MON and RON of three grades of gasoline

from the MONs and RONs ofthe gasoline blending stocks. It is well known that the

octane blending is not linear. In other words, the octane number ofthe gasoline is not

equal to the volumetric average ofthe octane numbers ofthe blending stocks. One way to

account for the nonlinearity is using interaction parameters. Morris et al. (1975) proposed

an equation for binary blending as below:

a = Xjaj +x^a,+b,2X,X2, (7.6)

where

a- RON or MON of a blend,

ai, a2- RON or MON ofthe blending component 1 and 2, respectively,

bi2- binary interaction parameter between component 1 and 2,

xi, X2- the volume fractions ofthe blending component 1 and 2, respectively.

The disadvantage of Morris' approach is that the interaction parameter between

each pair of gasoline cut has to be known for the blending. These interaction parameters

can only be obtained from the regression of industrial blending data. For example, in the

refinery considered in this work, blending of seven stocks needs twenty-one binary

parameters for MON and RON calculation each. Twu and Coon (1996) proposed a

method which only needs three interaction parameters for RON and MON each for

blending any number of gasoline blending stocks. Twu and Coon's approach uses the

concept of interaction parameter and also the compositions of gasoline cuts. It was

148

Page 164: 5-95 gap chart

claimed that the correlation describes blending behavior accurately throughout the entire

composition and it works not only for binaries, but also for multicomponent systems

(Twu and Coon, 1996). The average absolute deviation ofthe model predictions is 1.00%

for RON and 1.19% for MON in a gasoline blending test on a total of 161 blends from

157 gasoline cuts (Twu and Coon, 1996). Because of its simplicity and accuracy, the

approach of Twu and Coon (1996) has been used in this work.

In Twu and Coon's method, each gasoline blending stock consists three

components: aromatics, olefins, and saturates. The saturates include paraffins and

naphthenes. The approach uses several assumptions stated below:

Each ofthe three components in a blending stock has the same octane number as

the blending stock. For example, if the RON ofthe FCC gasoline is 87, then the RONs of

the olefins, the aromatics and the saturates in the FCC gasoline are all assumed to be 87.

a. The values ofthe interaction parameters between components inside the same

gasoline blending stock are zero.

b. If i and j are both the same hydrocarbon type (e.g., aromatics in cut A and aromatics

in cut B), the binary interaction parameter between i and j is assumed to be zero.

c. Gasoline blending is symmetric blending. That is, the binary interaction parameters

between two gasoline components i and j , kij and kji, have the same value.

d. The binary interaction parameters in the pairs with same components are the same.

For example, the binary interaction parameter between aromatics in cut A and olefins

in cut B is equal to that between the olefins in cut A and the aromatics in cut B.

The total number ofthe binary interaction parameters for blending two 3-

component gasoline cuts is only three (i.e., binary interaction parameters between

aromatics A and olefins in B, aromatics in A and saturates in B, olefins in A and saturates

in B). The binary interaction parameters between components are derived from

regressing blending data of gasoline blending stocks. The universal set of binary

interaction parameters between components were given by Twu and Coon (1996). They

were obtained from the regression of 161 blends from 157 gasoline cuts. The values of

the universal set of binary interaction parameters are listed in Table 7.6. Subscripts O, A,

149

Page 165: 5-95 gap chart

and S represent olefins, aromatics and saturates, respectively. For example, KQA

represents the binary interaction parameters between olefins and aromatics.

Table 7.6 Universal Set ofthe Binary Interaction Parameter between Components.

Interaction Parameter

K Q A

K Q S

K-AS

Twn and Coon

RON

0.0670

-0.1021

-0.0232

MON

0.0354

-0.0800

0.0271

Model Regression

RON

-0.00274

-0.1439

0.1841

MON

0.07014

-0.1385

0.1066

The binary interaction parameters between two gasoline blending stocks can be

calculated from the binary interaction parameters between components in these two

gasoline blending stocks. Assume that gasoline blending stock X is a mixture of three

components with compositions xi, X2, and X3, and the gasoline blending stock Y is a

mixture ofthe same three components with compositions yi, y2, and y^. For two three-

component gasoline blending stocks, the binary interaction parameter between gasoline

blending stocks X and Y is calculated using the following formula:

3 3

YJl(^>yj+^jy'^ K^ -1 I J

-^(^i+^j) (1-k,)

(a^+aj) (7.7)

where

ai- RON or MON ofthe component i,

kij- binary interaction parameter between the component i and j ,

KxY- binary interaction parameter between the gasoline blending stocks X and Y,

Xi, yi- volume fraction ofthe component i in gasoline blending stocks X and Y,

respectively.

150

Page 166: 5-95 gap chart

After generating the binary interaction parameters between gasoline blending

stocks, the octane numbers ofthe a gasoline blend with several gasoline blending stocks

are calculated using the formula given below:

where

Zi,Zj- volume fractions of gasoline blending stocks i and j , respectively,

a- RON or MON ofthe gasoline,

aij- the blending parameter.

The aij is calculated using the formula below:

ci,j=^(a,+aj)(l-k,j), (7.9)

where

ai, aj- RON or MON ofthe gasoline blending stocks i and j , respectively,

kij- binary interaction parameter between the gasoline blending stocks i and j .

Ethanol is very different from the other gasoline blending stocks. The ethanol

does not belong to any ofthe three molecule types of saturates, aromatics, and olefins.

Hence, the ethanol is treated different from the other seven gasoline blending stocks in

the octane model. Instead of using the real octane numbers ofthe ethanol, the blending

octane numbers are used for the ethanol. The ethanol only exists in the gasoline grade of

sub. octane with ethanol. The ethanol in this grade is fixed at IO.vol% according to

practice in the refinery considered in this work. After obtaining the octane number ofthe

gasoline blend of other seven blending stocks, the final octane number ofthe sub. octane

with ethanol gasoline is calculated as the volume average ofthe blending octane number

ofthe ethanol and the octane number ofthe gasoline blends of other seven gasoline

blending stocks.

151

Page 167: 5-95 gap chart

In order for the octane model to match the blending in the refinery considered in

this work, the three interaction parameters for RON and the three interaction parameters

for MON are regressed using the industrial blending data. The industrial blending data

used in the regression include 11 super unleaded blends, 67 unleaded blends, and 17 sub.

octane with ethanol blends. The properties of gasoline blending stocks in each blend are

unknown. Therefore, the average properties ofthe industrial data are used for the

gasoline blending stocks in each blend. These blends include both the blends in the

summer and in the winter. The regression is an optimization problem. A least square error

type fimction is used as the objective function for the optimization problem given below:

^W parameterization ~ 2^^ ^model.i ~ ^base.i ) ' ^base.i I ' ( ' • ' ^ j

where

Objfparamterization- objcctivc fiinction of the octane model parameterization,

Omodei.i- MON or RON predicted by the octane model,

Obase,i- MON or RON in the industrial data.

NPSOL optimization package (Gill et al., 1986) was used in the parameterization.

The average relative difference between industrial data and model prediction is 0.67% for

RON and it is 0.72% for MON. The maximum relative difference is 2.3% for RON and it

is 2.7% for MON. Considering the use of average properties for gasoline blending stocks

in each blend and different blends in the regression, the gasoline blending correlations

predict MON and RON accurately using only three interaction parameters each. The

interaction parameters obtained in the regression are listed in Table 7.6.

7.3.2 RVP Blending Model

Reid vapor pressure (RVP) is widely used as a criterion to measure the volatility

of gasoline. Theoretical approaches (Stewart, 1959; Vazquez-Esparragoza et al., 1992)

152

Page 168: 5-95 gap chart

use pseudo-components to represent the gasoline blending stocks. Empirical correlations

are used in the theoretical approaches to calculate the properties of pseudo-components,

which deteriorate the accuracy ofthe theoretical approaches. In addition, the

computations required in the theoretical methods are complex in comparison to those

required in other approaches (Singh et al., 2000). Chevron's vapor pressure blending

indices (VPBI) method uses empirical correlations. VPBI method has a simple form and

has been proved to have reasonable accuracy (Gary and Handwerk, 1984). It has been

widely used in industry (Singh et al., 2000). In the present work. VPBI method is used in

the RVP calculation.

The VPBI method converts the RVP of each gasoline blending stock to the

VPBI. The VPBI ofthe gasoline blend is calculated as the volumetric average ofthe

VPBIs of all the gasoline blending stocks. The RVP ofthe gasoline blend is obtained by

converting the calculated VPBI ofthe gasoline blend to RVP.

The relation between the VPBI and the RVP of a gasoline cut is given below:

VPBI = RVP"\ (7.11)

where

VPBI- vapor pressure blending index,

RVP- Reid vapor pressure.

Equation 7.11 can be modified to calculate RVP from VPBI.

153

Page 169: 5-95 gap chart

Start

Boiling points of butane, toluene, and

ethanol

Initial Guesses ofthe

T

Calculate the volume distilled for the mixed

blending stocks using Cubic Spline method

ASTM D86 curves of mixed blending

stocks

Calculate the total volume distilled from the mixed

blend stocks

Add the volume of pure component i to the total volume distilled, i=i+l

Calculate the new temperature using the

Secant method

J^Q_

Figure 7.1 Flowchart for the Calculation ofthe Tio»/„, T5o»/o, T9oo/„ of Gasoline Blends.

154

Page 170: 5-95 gap chart

7.3.3 Percent Distilled Model

The percent distilled model is to calculate the temperatures at 10%), 30%) and 90%

distilled ofthe gasoline based on the ASTM D86 curves ofthe gasoline blending stocks.

It is assumed that the gasoline is an ideal mixture in this work. The temperatures at 10%,

30% and 90%o distilled are represented by Tio%, or T5o%, or T9o%, respectively. The flow

chart for the above iterative procedure is shown in Figure 7.1.

The calculation of Tio%, or T5o%, or Tgoo/, follows the steps given below:

a. Make an initial guess of Tio%, or T5o%, or T9oo/„. Use the estimated value and the

ASTM D86 curves ofthe mixed blending stocks to calculate the volume percentages

ofthe mixed blending stocks distilled. Nine intermediate points are used to represent

each ASTM D86 curve. Cubic spline method (Riggs, 1994) is used to interpolate

between the known points on the ASTM D86 curves.

b. Calculate the total volume distilled by summing the volume distilled of each mixed

blending stocks.

c. Compare the temperature with the boiling points of three pure blending components:

butane, toluene, and ethanol. If the temperature is higher than one ofthe boiling

points, the corresponding component is assumed completely distilled. Then the

volume of this component is added to the total volume distilled. Repeat the procedure

until all three boiling points have been compared with.

d. Calculate the volume percentage that is vaporized. Compare the volume percentage

with 10 vol.%, 50 vol.%, or 90 vol.%). If they do not match, calculate the new

temperature using Secant method (Riggs, 1994) and return to step a.

155

Page 171: 5-95 gap chart

CHAPTER 8

REFINERY-WIDE OPTIMIZATION

In the previous chapters, the single-unit models in the overall refinery-wide model

and the feed characterizations that connect the single-unit models are presented. Till now,

the overall refinery-wide model has been built. This chapter discusses the optimization

analysis that was conducted on this refinery-wide model.

8.1 Formulation ofthe Optimization Problem

8.1.1 Formulation of the Obj ective Function

Although optimization can be stated in many different ways, the common

optimization to an industrial process is to maximize the profitability ofthe process. In

this work, the whole refinery is considered to be one process. This process uses the given

cmde slate to produce various petroleum products to achieve economic objectives.

The objective of optimization in hand is to reach the maximal profitability given

the cmde slate and refining facilities. No major hardware change in the current facilities

is considered in the optimization. The optimization tries to find the optimal operating

conditions that maximize the overall profit ofthe whole refinery while observing all the

process constraints.

In the objective fimction, only the feed cost, operating cost and revenue from

product sales are taken into account. The fixed costs, such as the capital cost and salaries,

are excluded from the objective fimction, since we can not influence these costs by

optimizing the operating variables.

The feed cost includes the cost of cmde oil feeds and the cost of other chemicals

purchased from the market. Four types of cmde oils are used in the refinery. These four

types of crude oils have been discussed in Chapter 3. It is assumed that each type of

crude oil has a fixed fraction for each operation mode. The compositions ofthe cmde

feed are shown in Table 3.1 of Chapter 3. The composition ofthe cmde oil for each

operation mode is fixed. Hence, a single cmde price is calculated from the prices of four

156

Page 172: 5-95 gap chart

types of crude oils and this single crude price is used as constant in each operation mode.

The chemicals purchased from the market and their usage are listed in Table 8.1.

Table 8.1 Chemicals Purchased by the Refinery Considered in This Work.

Chemicals

Butane

Isobutane

Toluene

Ethanol

Usage

Gasoline blending stock

Alkylation feed

Gasoline blending stock

Gasoline blending stock

The operating cost includes the costs in four categories: catalyst and additives,

natural gas, power, and cooling water. To calculate the exact usage of each category in

every unit in the refinery requires very detailed models for each unit, which is not in the

scope ofthe present work. Hence, all four categories are combined into one category

called utility. Since the operating costs are much smaller compared to other items in the

objective function, this simplification has an insignificant effect on the optimization

results. The utility cost is calculated based on the data in the linear programming (LP)

reports obtained from the refinery considered in this work. The utility costs are calculated

using the formula given below:

Utility i = — ^ • Utility i^Lp, (8.1) ^i,LP

where

Utilityi- utility cost for unit i,

Utilityi„LP- utility cost for unit i from LP report,

Vi- throughput of unit i,

Vi Lp- throughput of unit i from LP report.

157

Page 173: 5-95 gap chart

The income from the product sale includes the sales of all products produced in

the refinery. These products are listed in Table 8.2.

Table 8.2 Price Stmcture ofthe Refinery-Wide Optimization.

Name

Cmde

Butane (purchase)

Isobutane

Toluene

Ethanol

Propane

C3 product

Butane (sale)

Light Naphtha

Gasoline, Sub. Octane with ethanol, rack

Gasoline, Sub. Octane with ethanol, spot

Gasoline, unleaded, rack

Gasoline, unleaded, spot

Gasoline, super unleaded, rack

Gasoline, super unleaded.

Jet Fuel A

Jet Fuel 5

Low sulftir diesel

No. 2 diesel

No. 6 fuel oil

spot

Summer Mode

13.97

12.38

12.65

24.92

43.75

10.47

13.02

9.16

13.93

20.07

19.65

18.62

18.2

20.98

19.72

17.32

17.44

17.21

16.35

9.22

Winter Mode

14.54

14.55

11.33

23.47

43.29

11.85

13.96

11.33

14.60

18.90

18.48

17.17

16.75

18.30

17.04

17.01

17.28

16.65

16.04

13.2

158

Page 174: 5-95 gap chart

The objective function has the formula given below:

objf = Y Vproduct I • Ppruduct^ i

- 2_, VcrudCj - PcrudCj J

- 2 J Vchemical/^ - Pchemical,^ k

- Utility

- reformer regeneration cost (8.2)

where

Vproducti, VcmdCj, Vchemicalk- volumetric flow rate of product i, cmde type j , and

chemical k, respectively, BBL/Day,

Pproducti, Pcmdcj, Pchemicalk- price of product i, cmde type j , and chemical k,

respectively, $/BBL,

Utility- utility cost, $/Day,

Reformer regeneration cost- regeneration cost of reformer catalyst, $/Day,

It is clear that the objective function will be computed as $/day.

8.1.2 Objective Function Evaluation

Each objective function evaluation needs one execution ofthe re finery-wide

model to calculate the volumetric flow rates ofthe products. The utility cost and reformer

regeneration cost can only be evaluated through the execution ofthe refinery-wide

model.

The refinery-wide model essentially is a straight-through model. The sequence of

model executions follows the material flow in the refinery. During the refinery-wide

model execution, the cmde unit model is solved first. After that, the FCC model is

solved. Then the reformer model is solved twice, one execution for the low severity

159

Page 175: 5-95 gap chart

mode, the other execution for the high severity mode. The FCC model and the reformer

model can be solved simultaneously. The gas plant are executed after the executions of

the FCC model and the reformer model because the light gas products from the FCC unit

and reformer becomes the feed to the gas plant. The alkylation unit is then executed after

the execution ofthe gas plant model. Last, the gasoline blending model is executed. The

execution sequence ofthe refinery-wide model is shown in Figure 8.1. Although there are

recycle streams inside the single unit models, such as the hydrogen recycle in the

reformer unit, there is no recycle stream among the major processing units in the

refinery-wide model. In the refinery considered in this work, there are only a few streams

flowing in the direction opposite the execution sequence. However, these recycle streams

only have small quantities and are considered to have an insignificant effect on the

accuracy ofthe refinery-wide model.

8.1.3 Price Structure

The operation ofthe refinery considered in this work varies with seasons. This is

also tme for most of refineries in U.S. The operation changes are necessary to cope with

the market demands and price changes with respect to seasons. The price stmcture is

greatly affected by the market demand. For example, the gasoline has a higher price in

the summer when a lot of gasoline is consumed by travelers. The diesel has a higher price

in the winter because the demand of heating oil is high and diesel can be used to make the

heating oil.

The two major operation modes are Summer Mode and Winter Mode. Each

operation mode has its own price stmcture. The prices ofthe cmde oil, chemicals, and

products for two operation modes are listed in Table 8.2. It can be observed from Table

8.2 that the price gap between the gasoline and the diesel is large in the summer while the

same is small in the winter. This difference has great effect on refinery operation. Since

the price stmcture directly affects the value ofthe objective function of refinery-wide

optimization, optimization study is conducted for each mode separately.

160

Page 176: 5-95 gap chart

FCC unit Decision Variables

LSR

' '

Optimizer

Decision Variables 1 r Gasoline Blending

Refinery-wide Model

Interface

Decision Variables

Reformer Decision Variables

Crude Unit Decision Variables

Crude Uni t Model

FCC Model

FCC Gasoline

Lit 3ht Gas

Butane

1—•

' ' '

Crude Assays,

Crude Makeup

"

Reformer Model

(Low Severity, H igh Severity)

'

Gas Plant Model

' '

Alkylat ion Unit Model

' Alkylate

Gasoline Blending Model

Light Gas

Low Reformate, High Reformate

Figure 8.1 Flowchart ofthe Execution ofthe Refinery-Wide Model.

161

Page 177: 5-95 gap chart

8.1.4 Decision Variables

In this study, the decision variables are those variables whose values have

significant effect on the overall economics ofthe fuel-oriented refinery. These variables

are normally the handles that an operator can adjust to change the operation ofthe

refinery. For the point of view of modeling, the decision variables are the inputs to the

model whose optimal values are sought by the optimization algorithm. The number ofthe

decision variables is equal to the number of freedom degrees ofthe model used in the

optimization study. The decision variables for the refinery-wide optimization are the

collection ofthe decision variables of each unit in the refinery.

Based on the characteristics ofthe decision variables, the decision variables can

be divided into two groups: the decision variables ofthe processing units and the decision

variables ofthe gasoline blending. The decision variables ofthe processing units include

those process Vciriables, such as cut points, temperatures, recycle ratios, etc.

The gasoline blending is different from other units ofthe refinery in the sense that

it is only a mixing process. There is no chemical reaction or separation involved in the

gasoline blending. The decision variables for the gasoline blending are the fractions of

each gasoline blending stocks blended into three grades of gasoline products. Among the

seven blending stocks in the gasoline blending, five geisoline blending stocks, FCC

gasoline. Low severity reformate, high severity reformate, alkylate. Light straight-run

(LSR) gasoline, come from the processing units inside the refinery. In other words, there

is no purchase for these blending stocks from outside resources. Hence, the sum ofthe

fractions of each gasoline blending stock in three grades of gasoline is equal to unity. In

order to eliminate this equality constraint, only two fractions, instead of three, for each

blending stocks among these five blending stocks are used as the decision variables for

the refinery-wide optimization. The two fractions for each blending stock are the

fractions ofthe blending stock in the Super Unleaded Gasoline and Unleaded Gasoline.

The fraction ofthe blending stock in the Sub. Octane Gasoline, which now becomes a

dependent variable, is calculated using the formula given below:

162

Page 178: 5-95 gap chart

•^Sub.0ct.i •' ^ Super Unleaded.! -^Unleaded.!' \^-^)

where

Xsuper unieaded.i, Xunieaded.i, Xsub.Oct.i- fractions of blending stock i in Supct Unleaded,

Unleaded, and Sub. Octane Gasoline, respectively.

The refinery usually purchases some butane from outside resources in the winter

and sells some butane on the spot market in the summer. This is because of different

evaporation specifications for gasoline in different seasons. Gasoline in the winter has a

higher RVP upper limit so that the refinery can purchase butane and blend it into the

gasoline. Since butane has a lower market price than gasoline, it is beneficial to blend as

much as butane into the gasoline. The refinery sells part of butane produced inside the

refinery because the RVP upper limit is low and the amount of butane used in the

gasoline blending is low.

It is assumed initially that the total amount of butane used in the blending in the

summer is equal to what is produced within the refinery. It is assumed initially that the

total amount of butane used in the blending in the winter is equal to the sum ofthe butane

produced plus five thousand barrels per day purchase of butane. The five thousand barrels

per day purchase of butane is the maximum amount of butane that refinery usually

purchases from outside resources. In both modes, the tme amount of butane used in the

gasoline blending may not be equal to amount assumed initially. The assumed amount is

used as the base to calculate the amount of butane in barrel per day used in three grades

of gasoline from the fractions, which are the decision variables for the refinery-wide

optimization. Since the total amount of butane used in the gasoline blending is unknown

beforehand in the refinery-wide model, the sum ofthe fractions of butane in three grades

of gasoline is not necessarily to be unity. Hence, three fractions must be used for butane.

The toluene used in the blending comes only from outside resources since the

refinery does not produce toluene. The maximum amount of toluene used in the blending

is set at 200 barrels as per the operation department in the refinery. The 200 barrels per

day is used to calculate the amount of toluene in barrel per day from the fractions in the

163

Page 179: 5-95 gap chart

decision variable set. Three fractions for toluene are also used since the real amount of

toluene used in the blending is unknown beforehand.

One special decision variable in the gasoline blending is the fraction of Light

Straight-run (LSR) gasoline for sale. The refinery considered in this work sells part of

LSR on the spot market. LSR can be either sold or blended into the gasoline.

The decision variables and their lower and upper bounds are listed in Table 8.3.

The lower and upper bounds ofthe decision variables came from the normal operating

range used in the refinery considered in this work. The engineers from the refinery have

agreed upon the values of these lower and upper bounds.

Table 8.3 Decision Variables of Refinery-Wide Optimization.

Decision variable

No. Decision variable name

3

4

5

Cmde Unit: 1 Feed to cmde unit

2 Atmospheric Tower furnace outlet

temperature, °F

Light naphtha ASTM 95% point, °F

Heavy naphtha ASTM 95% point, °F

Jet Fuel ASTM 95% point, °F

6 Diesel ASTM 95% point, °F

7 Heavy Vacuum Gas Oil TBP end point, °F

FCC Unit: 8 Regenerator temperature, °F

9 Reactor temperature, °F

10 02% in the flue gas, mol%

Reformer: 11 First bed inlet temperature, K, low severity

12 H2/Hydrocarbon recycle ratio, mol H2/mol

Hydrocarbon, low severity

13 First bed inlet temperature, K, high severity

Lower Bound

Upper Bound

35,000

650

230

360

460

620

950

1,250

980.00

0.01

900.00

5.00

50,000

665

260

385

525

670

1,100

1,340

1,005

0.02

980.00

7.50

900.0 980.0

164

Page 180: 5-95 gap chart

Table 8.3 Continued.

Decision variable

No. Decision variable name Lower Bound

Upper Bound

Reformer: 14 H2/Hydrocarbon recycle ratio, mol Hi/mol 5.0

Hydrocarbon, high severity

15 Low severity operation time/total operation 0.1

time

16 Fractionof FCC gasoline in Super Unleaded 0.0

17 Fraction of low reformate in Super Unleaded 0.0

Blending:

Super

Unleaded

Unleaded

Sub. Oct.

Light

Naphtha:

18 Fraction of alkylate in Super Unleaded

19 Fraction of LSR in Super Unleaded

20 Fraction of high reformate in Super Unleaded 0.0

21 Fraction of butane in Super Unleaded

22 Fraction of toluene in Super Unleaded

23 Fraction of FCC gasoline in Unleaded

24 Fraction of low reformate in Unleaded

25 Fraction of alkylate in Unleaded

26 Fraction of LSR in Unleaded

27 Fraction of high reformate in Unleaded

28 Fraction of butane in Unleaded

29 Fractionof toluene in Unleaded

30 Fraction of butane in Sub. Oct.

31 Fraction of Toluene in Sub. Oct.

32 Fraction of Light naphtha sold directly

7.5

0.80

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

1.0

1.0

1.0

1.0

1.0

1.0

I.O

I.O

1.0

I.O

1.0

I.O

I.O

1.0

1.0

1.0

1.0

165

Page 181: 5-95 gap chart

8.1.5 Constraints

There are many constraints in the operation of a fuel-oriented refinery. The

optimization problem in this study includes 63 nonlinear constraints and 7 linear

constraints. In this study, the nonlinear constraints are not expressed explicitly. Instead,

all the nonlinear constraints are evaluated using the nonlinear refinery-wide model.

These constraints can be divided into several categories: capacity limits, process variable

constraints, and product specifications.

A capacity limit is the maximum throughput of a unit. The capacity limits ofthe

refinery-wide optimization were obtained from the Linear Programming (LP) reports

obtained from the refinery considered in this work.

Process variable constraints are the maximum or minimum value of a process

variable. If the value of a process variable is outside the interval defined by its maximum

and minimum values, violation of safety regulation or damage to the facility may happen.

For instance, a temperature can not surpass the metallurgical limit ofthe contacting

materials.

The product specifications in this study are mainly the specifications for the three

grades of gasoline products: Super Unleaded Gasoline, Unleaded Gasoline, and Sub.

Octane Gasoline with Ethanol. The gasoline specifications from Designation D4814-96

of ASTM standards are used in this study. It includes eleven specifications for each grade

of gasoline. It should be noted that there are different product specifications for gasoline

vaporization in different seasons.

The nonlinear constraints are listed in Table 8.4. Engineers in the refinery

considered in this work have agreed upon the values ofthe upper limits and the lower

limits of these nonlinear constraints.

166

Page 182: 5-95 gap chart

Table 8.4 Nonlinear Constraints of Refinery-wide Optimization.

Nonlinear Constraint

Crude Unit:

FCC Unit:

No

I

2

3

4

5

6

7

8

9

10

11

12

13

. Constraint name

Diesel flow rate in

Atmospheric tower

Summer Mode

Lower Bound

0

Reduced cmde to vacuum 4,000

tower

Gasoline splitter bottom

flow rate

Crude debutanizer

Rose unit capacity

Rerun unit capacity

Naphtha hydrotreater

capacity

Jet fuel tank capacity

Heater outlet temperature,

op

Fuel flow to furnace.

SCM

Reactor pressure, psig

Regenerator pressure,

psig

Actual speed of lift air

blower (sa), RPM

0

0

0

0

0

0

500

0

14.7

14.7

5,100

Upper Bound

12,000

23,000

1,000

10,000

6,500

5,000

10,000

5,800

700

39.5

30

30

6,100

Winter Mode

Lower Bound

0

4000

0

0

0

0

0

0

500

0

14.7

14.7

5,100

Upper Bound

12,000

23,000

1,000

10,000

6,500

5,000

10,000

5,800

700

39.5

30

30

6,100

167

Page 183: 5-95 gap chart

Table 8.4 Continued.

Nonlinear No. Constraint name Constraint

Summer Mode Winter Mode

Lower Bound

Upper Bound

Lower Bound

Upper Bound

14

15

16

17

18

19

20

21

22

Difference between lift-aii

blower inlet suction flow

and surge flow, ICFM

Regenerator reactor

pressure difference (P6-

P4), psia

Level of catalyst in the

standpipe. foot

Combustion air blower

inlet suction flow, SCFM

Wet gas compressor inlet

suction flow, Ibmol/s

Combustion air blower

flow rate (Ib/s)

FCCU charge

Reformer Charge

Research Octane of Low

•0

-5

0

35,000

0

30

0

6,000

94

Infinite

2

20

42,000

0.67

65

20,000

10,000

96.5

0

-5

0

35,000

0

30

0

6,000

94

Infinite

2

20

42,000

0.67

65

20,000

10,000

96.5

reformate

23 Research Octane of High 98.5

reformate

Other 24 Distillate HDS 0

constraints: 25 Alkylation unit charge 0

26 Alkylate product 0

100.5 98.5 100.5

14,500

8,000

4,000

0

0

0

14,500

8,000

4,000

168

Page 184: 5-95 gap chart

Table 8.4 Continued.

Nonlinear Constraint

Blending:

Super

Unleaded

Gasoline

Unleaded

Gasoline

No.

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

Constraint name

C3 splitter

MON

(R+M)/2

RVP

Specific gravity

Sulftu-

Drivability index

10%) vaporized

50%) vaporized

90% vaporized

02 wt.%

Benzene, vft.Vo

MON

(R+M)/2

RVP

Specific gravity

Sulftir

Drivability index

10% vaporized

50%o vaporized

90% vaporized

02 wt.%

Summer Mode

Lower Bound

0

83

93.1

6.4

0.72

0

0

0

170

0

0

0

82.1

87.1

6.4

0.72

0

0

0

170

0

0

Upper Bound

15,000

Infinite

94

7.8

0.784

O.I

1250

158

250

374

4

4.9

84

88.1

7.8

0.784

0.1

1250

158

250

374

4

Winter Mode

Lower Bound

0

83

93.1

6.4

0.72

0

0

0

170

0

0

0

82.1

87.1

6.4

0.72

0

0

0

170

0

0

Upper Bound

15,000

infinite

94

12.5

0.784

O.I

1200

131

235

365

4

4.9

84

88.1

12.5

0.784

O.I

1200

131

235

365

4

169

Page 185: 5-95 gap chart

Table 8.4 Continued.

Nonlinear Constraint

Sub. Oct.

No.

49

50

with ethanol 51

Gas

Contract

52

53

54

55

56

57

58

59

60

61

62

63

Constraint name

Benzene, wt.%

MON

(R+M)/2

RVP

Specific gravity

Sulfiir

Drivability index

10%) vaporized

50% vaporized

90% vaporized

02 wt.%

Benzene, wt.%

Super Unleaded Rack

Unleaded Rack

Sub. Oct. with Ethanol

Summer Mode

Lower Bound

0

82.1

87.1

6.4

0.72

0

0

0

170

0

3.7

0

3300

16000

3700

Upper Bound

4.9

84

88

8.8

0.784

0.1

1250

158

250

374

4

4.9

3300

16000

3700

Winter Mode

Lower Bound

0

82.1

87.1

6.4

0.72

0

0

0

170

0

3.7

0

3300

16000

3700

Upper Bound

4.9

84

88

13.5

0.784

0.1

1200

131

235

365

4

4.9

3300

16000

3700

Linear constraints are only for the fractions ofthe gasoline blending stocks in the

three grades of gasoline blending. They have a common formula given below:

l>Y^j,>0, (8.4)

where

j - Gasoline blending stocks.

170

Page 186: 5-95 gap chart

i - grades of gasoline,

Xji- fraction of gasoline blending stock j in the gasoline of grade i.

Since there are seven gasoline blending stocks, there are seven linear constraints.

It should be noted that for the linear constraints corresponding to those five blending

stocks only produced inside the refinery, only two fractions are used in each linear

constraint.

8.1.6 Optimization Algorithm

Each objective fimction evaluation requires solving the refinery-wide model once.

Solving the refinery-wide model is time consuming since the refinery-wide model is

fairly large. Therefore, it is necessary to find an optimization algorithm that has fast

convergence that requires less objective function evaluations. It is reported that

successive quadratic programming (SQP) converges fast and it may be the best method in

solving nonlinear problem (Edgar and Himmelblau, 1988). A software package using

SQP algorithm, known commercially as NPSOL was used as the optimization engine.

The documentation of NPSOL software was given by Gill et al. (1986).

From a random starting point, NPSOL can not guarantee the convergence to the

global optimum, nor does any other software package. The best that the NPSOL can do is

to converge to a point that satisfies the first-order Kuhn-Tuck conditions, which is the

necessary conditions for a local minimum. To increase the possibility of finding the

global optimum, several random starting points are used for each case discussed below.

The solution with the best value ofthe objective function is regarded as the global

optimum and is used in the following analysis.

8.2 Optimization Case Studies

8.2.1 Base Case

In the refinery considered in this work, unit operations follow the schedule

determined by the linear programming (LP). However, the decision variables used by LP

are not exactly the same as those used in nonlinear optimization studies. It is necessary to

171

Page 187: 5-95 gap chart

know the normal operating values ofthe decision variables used in optimization studies

in order for fair comparison.

The normal values of some decision variables are obtained from normal operating

conditions in the refinery. The normal values ofthe rest ofthe decision variables are

found by trial and error to make the model prediction consistent with the product

information in the LP reports. The complete sets ofthe normal values ofthe decision

variables for both modes are listed in Table 8.6 and Table 8.9. Engineers from the

refinery have agreed upon these normal values. These two sets of normal values are

regarded as the base cases in this work.

It is desirable to mn the refinery-wide model using these normal operation

conditions and compare the product slate and purchase plan with the LP reports. The

results are listed in Table 8.5 and Table 8.6.

It can be seen from Table 8.5 and Table 8.6 that the average relative errors ofthe

volumetric flow rates of products and purchases ofthe LP reports and model predictions

under normal operating conditions are 1.36% for Summer Mode and 1.31% for Winter

Mode. There are relative big difference for the quantitis of propane, ethanol, butane and

C3 product in both modes or one mode. However, those streams have relative small

amount. Therefore, the mismatch have an insignificant effect on the overall revenue of

the refinery.

8.2.2 Summer Mode

The refinery-wide optimization was conducted for Summer Mode. The optimum

values ofthe decision variables of refinery-wide optimization are listed in Table 8.7. The

active constraints in the optimum solution are listed in Table 8.8. The comparison ofthe

product slates between the optimum solution and base case is shown in Table 8.9 and

Figure 8.2.

172

Page 188: 5-95 gap chart

Table 8.5 Comparison of Model Prediction ofthe Base Case with the LP report. Summer Mode.

Item

Purchase, bbl/day

Cmde

Toluene

Ethanol

Isobutane

Sale, bbl/day

Light naphtha

Butane

SNL spot

NL spot

Sub.Oct. spot

SNL rack

NL rack

Sub. Oct. rack

Propane

C3 product

Jet fiiel A

Jet fiiel 5

Low sulfur diesel

No. 2 diesel

No.6 fiiel oil

Average error

Purchase, $/day

Sale, $/day

Utility, $/day

Revenue, $/day

LP report

50000

200

379

1377

1017

950

0

1742

0

3300

16000

3700

371

1254

800

5000

10813

3419

2440

-

737529

881742

24671

119543

Model Prediction

50000

198

370

1,407.4

1009.5

927.17

0

1744.5

0

3300

16000

3700

311

1264

800

5000

10812

3418

2460

-

737470

881121

23897

119754

Relative Deviation, %

0.00

-1.00

-2.37

2.21

-0.74

-2.40

0.00

0.14

0.00

0.00

0.00

0.00

-16.10

0.80

0.00

0.00

-0.01

-0.03

0.82

1.36

-0.01

-0.07

-3.14

0.18

173

Page 189: 5-95 gap chart

Table 8.6 Comparison of Model Prediction ofthe Base Case with the LP Report, Winter Mode.

Item

Purchase, bbl/day

Cmde

Toluene

Ethanol

Butane

Isobutane

Sale, bbl/day

Light naphtha

SNL spot

NL spot

Sub.Oct. spot

SNL rack

NL rack

Sub. Oct. rack

Propane

C3 product

Jet fuel A

Jet fiiel 5

Low sulfur diesel

No. 2 diesel

No.6 fiiel oil

Average Error

Purchase, $/day

Sale, $/day

Utility, $/day

Revenue, $/day

LP report

50000.000

18.000

362.000

1044.300

945.000

595.000

0.000

3853.290

0.000

3300.000

16000.000

3700.000

129.000

1700.000

800.000

5000.000

13024.000

1339.000

1823.000

-

771508.105

865929.018

22843.928

71576.984

Model Prediction

50000.000

18.000

370.000

1041.630

945.440

600.450

0.000

3839.810

0.000

3300.000

16000.000

3700.000

108.020

1754.770

800.210

5000.000

13000.000

1321.680

1819.520

771821.816

865578.997

21932.920

71824.261

Relative Deviation, %

0.00

0.00

2.21

-0.26

0.05

0.92

0.00

-0.35

0.00

0.00

0.00

0.00

-16.26

3.22

0.03

0.00

-0.18

-1.29

-0.19

1.31

0.04

-0.04

-3.99

0.35

174

Page 190: 5-95 gap chart

It should be noted that the revenue in the optimum solution shown in Table 8.8

does not include the reformer regeneration cost. The reason is that the reformer operation

is not stable in the refinery considered in this work and the cycle length varies in the real

operation. It is difficult to estimate the regeneration cost per day for the base case. Hence,

the reformer regeneration cost is not included in the revenue ofthe base case.

To make a fair comparison, the reformer regeneration cost is also not included in

the revenue of the optimum solution. Same approach is used in the case of Winter Mode

and the case of single unit optimization.

It should be noted that the cmde throughput in both the optimum solution and the

base case reached the upper limit ofthe cmde unit capacity. Hence, no benefit is obtained

from incremental throughput in the optimum solution. The only thing the optimizer can

do to improve the profitability ofthe refinery is to adjust the product slate and make more

high-value products and less low-value products.

It can be seen from Figure 8.2 that the significant differences between base case

and the optimum solution are the gasoline production and diesel production. It should be

noted that the refinery has contracts with retailing companies for all three grades of

gasoline. These contracts become the lower limits for gasoline production. In other

words, the refinery must produce certain amount for each grade of gasoline. Any amount

of gasoline above the amount in the contract is sold on the spot market. In the optimum

solution, more spot Unleaded Gasoline is produced them the base case. This is

understandable because the Unleaded Gasoline has a relative high price in Summer Mode

The question may be raised that why the optimizer does not try to make more spot

Super Unleaded Gasoline which has a higher price than the Unleaded Gasoline. The

answer is that to make more Super Unleaded Gasoline, either the reformer will be

operated longer at high severity mode, or the FCC unit will be operated at higher

severity, or more toluene is purchased. None ofthe three options is favorable from the

economic point of view. Operating the reformer longer at high severity mode decreases

the conversion from the feed to reformate and produces more low-value light gas. It also

shortens the cycle length ofthe reformer, which results in decrease of on-stream factor

175

Page 191: 5-95 gap chart

and increase ofthe regeneration cost. For FCC operation, since the FCC operation has

already reached the upper limit ofthe wet gas compressor constraint in the optimum

solution, decreasing throughput is necessary when the severity of FCC unit is increased.

Increasing the severity of FCC unit also increases the production of low-value light gas.

Purchasing more toluene is also not a good option because the toluene price is higher

than that of gasoline. Therefore, it is a better choice to make more spot Unleaded

Gasoline instead of spot Super Unleaded Gasoline.

The extra spot Unleaded Gasoline produced in the optimum solution mainly

comes from cutting more diesel components into the gasoil and blending all light naphtha

in the gasoline. It is observed from Figure 8.2 that less diesel is produced in the optimum

solution. The light naphtha has a relative low market price and it should be all blended

into gasoline.

The optimum solution also shows that low severity operation time/total operation

time reaches the upper limit. It means the low severity operation should be longer than

normal operation. This can be explained by the fact that high severity operation produces

more low-value light gas and less high-value reformate, though the octane number ofthe

reformate is higher. High severity operation also results in more coke deposited on the

catalyst and consequently, shorter cycle length. The gain of operating reformer at high

severity longer is that more Super Unleaded Gasoline can be produced. However, it

seems that the cost of producing more Super Unleaded Gasoline is too high. The

information sent by the optimum solution is that as long as there are enough octane

number in the gasoline, the reformer should be operated at low severity in Summer

Mode.

176

Page 192: 5-95 gap chart

Table 8.7 Optimum Values ofthe Decision Variables of Refinery-wide optimization. Summer Mode.

Decision

variables

No. Decision variable name Base Case

Summer

Optimum

Summer

Crude Unit: I

2

3

4

5

6

7

Feed to crude unit

AT furnace outlet temperature, °F

Light naphtha ASTM 95% point, °F

Heavy naphtha ASTM 95% point, °F

Jet Fuel ASTM 95% point, °F

Diesel ASTM 95% point, °F

Heavy Vacuum Gas Oil TBP end point, °F

FCC Unit: 8 Regenerator temperature, °F

9 Reactor temperature, °F

10 02%) in the flue gas, mol%

Reformer: 11 First bed inlet temperature, K, low severity

12 H2/Hydrocarbon recycle ratio, mol/mol, low

severity

13 First bed inlet temperature, K, high severity

14 H2/Hydrocarbon recycle ratio, mol/mol, high

severity

15 Low severity operation time/total operation

time

16 Fraction of FCC gasoline in Super Unleaded

17 Fraction of low reformate in Super Unleaded

Blending

Super

Unleaded

50000.000

665.000

260.000

370.000

475.350

622.000

1091.000

1270.000

1000.000

0.013

900.000

6.000

912.000

5.500

50000.000

664.999

230.228

384.824

489.021

634.983

1099.460

1250.013

1002.251

0.020

900.000

7.486

911.736

7.309

0.750

18 Fraction of alkylate in Super Unleaded

19 Fraction of LSR in Super Unleaded

20 Fraction of high reformate in Super Unleaded 0.750

21 Fraction of butane in Super Unleaded

0.800

0.000

0.001

0.565

0.000

0.750

0.054

0.079

0.000

0.573

0.000

0.040

0.047

177

Page 193: 5-95 gap chart

Table 8.7 Continued.

Decision

variable

No. Decision variable name Base Case

Summer

0.990

0.991

0.705

0.176

0.321

O.I 00

0.187

0.000

0.034

0.000

0.363

Optimum

Summer

1.000

0.910

0.710

0.236

0.600

0.922

0.106

0.000

0.035

0.000

0.000

Unleaded

Sub. Oct.

Light

naphtha

22 Fraction of toluene in Super Unleaded

23 Fraction of FCC gasoline in Unleaded

24 Fraction of low reformate in Unleaded

25 Fraction of alkylate in Unleaded

26 Fraction of LSR in Unleaded

27 Fraction of high reformate in Unleaded

28 Fraction of butane in Unleaded

29 Fractionof toluene in Unleaded

30 Fraction of butane in Sub. Oct.

31 Fraction of Toluene in Sub. Oct.

32 Fraction of Light naphtha sold directly

Table 8.8 Active Constraints in Refinery-wide Optimization, Summer Mode.

Active Constraint Name Limit type

Jet Fuel Tank Capacity Upper

Wet gas compressor inlet suction flow Upper

FCC throughput Upper

Super Unleaded Gasoline, (R+M)/2 Lower

Super Unleaded Gasoline, RVP Upper

Super Unleaded Gasoline, Specify Gravity Lower

Unleaded Gasoline, MON Lower

Unleaded Gasoline, RVP Upper

Sub. Oct. Gasoline, (R+M)/2 Lower

Sub. Oct. Gasoline, RVP Upper

Sub. Oct. Gasoline volumetric flow rate Lower

178

Page 194: 5-95 gap chart

Table 8.9 Comparison ofthe Decision Variables of Refinery-wide Optimization with Base Case, Summer Mode.

Item

Purchase, bbl/day

Crude

Toluene

Ethanol

Isobutane

Sale, bbl/day

Light naphtha

Butane

SNL spot

NL spot

Sub.Oct. spot

SNL rack

NL rack

Sub. Oct. rack

Propane

C3 product

Jet ftiel A

Jet ftiel 5

Low sulftir diesel

No. 2 diesel

No.6 fuel oil

Summary

Purchase, $/day

Sale, $/day

Utility, $/day

Revenue, $/day

Model Prediction

50000

198

370

1,407.4

1009.5

927.17

0

1744.5

0

3300

16000

3700

311

1264

800

5000

I08I2

3418

2460

737470

881121

23897.00

119754

Optimum Solution

50000

200

370

1314

0

976

72

3009

0

3300

16000

3700

306

1192

800

5000

10539

3332

2464

736338.6

884888.93

23467.00

125083.33

Relative Deviation, %

0.00

I.Ol

0.00

-6.64

-100.00

5.27

-

72.48

-

0.00

0.00

0.00

-1.69

-5.70

0.00

0.00

-2.52

-2.51

0.17

-0.15

0.43

-1.80

4.45

179

Page 195: 5-95 gap chart

<L> xn CO

U

OS OQ

fe rt N

• ^H

s 4-» & I

o <u T3

^

i u

d)

c<

( N

o Z

% o

<n (L>

u

dies

U l

e. 3 l/J

_

(U

•o

o

ner

M

3 m VI cS

U ID

<u <s 1 .

<u • ~ >

u c es n. o I-I

u.

•^ o o

X3 3 on

O O . w

J

z o o. U5

J

z !/3

4) c BS

-^.» 3 X I

igh

-» o npo

I - I D .

t ' O o.

« J=

a. ta c

OQ

c o

N

O

ID

c

o CO

o 3 o

c o Vi

. I—I

53 O H £ o O (N 00 (U l i

3

O O O

O O o o

o o o 00

o o o so

o o o o o o

o/iee

180

Page 196: 5-95 gap chart

Some spot Super Unleaded Gasoline is produced in the optimum solution of

Summer Mode. This increase in Super Unleaded Gasoline is realized by blending

gasoline closer to its lower limit of octane specifications. In the gasoline blending process

ofthe refinery considered in this work, the octane number ofthe gasoline is usually

blended to about 0.1 to 0.4 octane number higher than its specifications. This is a

common industrial practice to avoid disqualification and reblending. Since the refinery

considered in this work has not implemented advanced control system, which is a

necessity for tight gasoline blending, the octane giveaway can not be decreased. In the

optimum solution, either one ofthe two octane specifications, (R0N+M0N)/2 or MON,

reaches the lower limit. By blending the gasoline closer to the octane specification, more

Super Unleaded Gasoline can be produced.

It is found that gasoline blending agrees with the common industrial practice by

hitting the lower limits of octane number specification and the upper limits of RVPs. The

upper limits of RVPs are always active because by doing so, more low-value butane can

be blended into high-value gasoline.

8.2.3 Winter Mode

The refinery-wide optimization was conducted for the Winter Mode. The

optimum operating conditions that maximize the revenue were found. The optimum

values ofthe decision variables of refinery-wide optimization are listed in Table 8.10.

The active constraints in the optimum solution of Winter Mode are listed in Table 8.11.

The comparison ofthe product slates ofthe optimum solution with the base case is shown

in Table 8.12 and Figure 8.3.

It can be seen from Figure 8.3 that the significant differences between base case

and optimum solution are also the gasoline production and diesel production.

In the optimum solution, the refinery produces significant amount of spot Super

Unleaded Gasoline where there is no spot Super Unleaded Gasoline in the base case. The

optimizer adjusts the operating conditions ofthe processing units and the gasoline

181

Page 197: 5-95 gap chart

blending to maximize the Super Unleaded Gasoline, which has the highest market value

among all the grades of gasoline.

More cmde has been cut into diesel in the optimum solution than in the base case.

This can be explained by the fact that diesel has high market value in the Winter Mode.

Although diesel price is still lower than the gasoline price in the winter in this case, there

is more volumetric lost in gasoline production route. The loss mainly happens in the FCC

unit and the reformer where large amount of low-value light gases is produced.

Considering both the price gap and the volumetric loss, it is more beneficial to produce

more diesel in Winter Mode.

However, there is also a limit of increasing diesel production. Increasing diesel

product means less feed for the FCC unit. Further decrease in FCC feed may result in

overcracking in FCC unit if the riser temperature remains high. Overcracking produces

more light gases, which result in more volumetric loss for gasoline production. If the riser

temperature is lowered to avoid overcracking, the octane number ofthe FCC gasoline

will drop, which results in less Super Unleaded Gasoline that can be produced since the

octane number of FCC gasoline drops. The optimizer makes a compromise between

diesel production and gasoline production and chooses an appropriate value for the

ASTM 95% point ofthe diesel that maximizes the profitability ofthe refinery.

Again, the light naphtha has a relative low market price. The optimum solution

shows that it should be all blended into gasoline. The optimum solution also shows that

low severity operation time/total operation time reaches the upper limit. It means the low

severity operation should be longer than normal operation in the Winter Mode. This can

be explained by the factor that the gasoline pool has a higher average octane number

because more high-octane butane has been blended into the gasoline pool. Hence, there is

less demand on the high octane reformate produced by high severity reformer operation.

182

Page 198: 5-95 gap chart

Table 8.10 Comparison ofthe Decision Variables of Refinery-wide Optimization with Base Case, Winter Mode.

Decision

variable

No. Decision variable name Base Case

Winter

Optimum

Winter

Feed to cmde unit

AT fiimace outlet temperature, °F

Light naphtha ASTM 95% point, °F

Heavy naphtha ASTM 95% point, °F

Jet Fuel ASTM 95% point, °F

Diesel ASTM 95% point, °F

Heavy Vacuum Gas Oil TBP end point, °F

8 Regenerator temperature, °F

9 Reactor temperature, °F

10 02% in the flue gas, mol%

11 First bed inlet temperature, K, low severity

12 H2/Hydrocarbon recycle ratio, mol/mol, low

severity

13 First bed inlet temperature, K, high severity

14 H2/Hydrocarbon recycle ratio, mol/mol, high

severity

15 Low severity operation time/total operation

time

16 Fraction of FCC gasoline in Super Unleaded

17 Fraction of low reformate in Super Unleaded

18 Fraction of alkylate in Super Unleaded

19 Fraction of LSR in Super Unleaded

20 Fraction of high reformate in Super Unleaded

21 Fraction of butane in Super Unleaded

Crude Unit:

FCC Unit:

Reformer:

Blending

Super

Unleaded

50000.000

665.000

250.000

370.000

465.280

593.000

1055.000

1270.000

1000.000

0.013

900.000

6.000

910.101

5.500

50000.000

650.000

246.669

364.123

460.000

620.000

1070.000

1300.531

1004.997

0.012

900.000

7.500

910.000

7.500

0.750 0.800

0.000

0.000

0.485

0.000

0.846

0.053

0.000

0.112

0.823

0.000

1.000

0.077

183

Page 199: 5-95 gap chart

Table 8.10 Continued.

Decision

variables

No. Decision variable name Base Case

1. Winter

0.090

0.900

0.999

0.267

0.526

0.025

0.209

0.000

0.045

0.000

0.224

Optimum

2. Winter

0.000

0.862

0.888

0.000

0.694

0.000

0.186

0.000

0.042

0.000

0.000

22 Fraction of toluene in Super Unleaded

23 Fraction of FCC gasoline in Unleaded

24 Fraction of low reformate in Unleaded

25 Fraction of alkylate in Unleaded

26 Fraction of LSR in Unleaded

27 Fraction of high reformate in Unleaded

28 Fraction of butane in Unleaded

29 Fraction of toluene in Unleaded

30 Fraction of butane in Sub. Oct.

31 Fraction of Toluene in Sub. Oct.

32 Fraction of Light naphtha sold directly

Table 8.11 Active Constraints in Refinery-wide Optimization, Winter Mode.

Active Constraint Name Limit type

Jet Fuel Tank Capacity Upper

Wet gas compressor inlet suction flow Upper

Super Unleaded Gasoline, (R+M)/2 Lower

Super Unleaded Gasoline, RVP Upper

Super Unleaded Gasoline, Specify Gravity Lower

Unleaded Gasoline, MON Lower

Unleaded Gasoline, RVP Upper

Sub. Oct. Gasoline, MON Lower

Sub. Oct. Gasoline, RVP Upper

Sub. Oct. Gasoline volumetric flow rate Lower

184

Page 200: 5-95 gap chart

Table 8.12 Comparison ofthe Product Slates ofthe Optimum Solution with the Base Case, Winter Mode.

Base Case Optimum Solution Relative Deviation, %>

Purchase, bbl/day

Crude

Toluene

Ethanol

Butane

Isobutane

Sale, bbl/day

Light naphtha

SNL spot

NL spot

Sub.Oct. spot

SNL rack

NL rack

Sub. Oct. rack

Propane

C3 product

Jet ftiel A

Jet ftiel 5

Low sulfiir diesel

No. 2 diesel

No.6 fuel oil

Summary

Purchase, $/day

Sale, $/day

Utility, $/day

Revenue, $/day

50000.000 50000

18.000 0

370.000 370

1041.630 1047.78

945.44 818.68

600.450

0.000

3839.810

0.000

3300.000

16000.000

3700.000

108.020

1754.770

800.210

5000.000

13000.000

1321.680

1819.520

0

1495.56

1656.24

0

3300.000

16000.000

3700.000

113.98

1579.99

800

5000

13000

2755.94

1806.76

771821.816 769691.3814

865578.997 866186.1954

21932.920 22097.12

71824.261 74397.694

0.00

-100.00

0.00

0.59

-13.41

0.92

-56.87

0.00

0.00

0.00

0.00

5.52

-9.96

-0.03

0.00

0.00

108.52

-0.70

-0.28

0.07

0.75

3.58

185

Page 201: 5-95 gap chart

(U Vi Ti U Vi

ca PQ D

3 O

efl N

s +.J O H

O (U

T 3

^ ^ U 3

"-! (^ •

<u cH vo o Z

CN

O Z

k« - 1

I * -

3 !/) » o —' m m

u , 3

o

" M V\ u

T3

_ M

0> r/1 W

TD

cS

U o

CS a. o u a.

^ CJ o

X 3 tn

O a. « J Z . -o D .

(/) J

z CA

.^.J

igh

~ ™

\ ' o o. CO

ca J3

Q. C3 C

O

I-I u 3

Vi ca U 0) c/3 ca

PQ

3 O

• ^

N

O,

O -a

3

c/5

;§ ^ - » o 3

T3 O

O 3 O Vi

'C ca o, B o

O rn 00

u

op

o o o 'a-

o o o cs

o o o o

o o o 00

o o o

o o o

o o o CN

o/iaa 186

Page 202: 5-95 gap chart

It is found from Table 8.11 that gasoline blending also agrees with the common

industrial practice by hitting the lower limits of octane number specification and the

upper limits of RVPs. The upper limit ofthe wet gas compressor is also an active

constraint. This means that even in the Winter Mode, the highest severity ofthe FCC unit

is favorable. This can be explained by the fact that the higher the FCC riser temperature

is, the higher the octane number ofthe FCC gasoline, and the more super unleaded

gasoline can be produced.

The wet gas flow rate is a function of FCC feed rate and riser temperature. The

optimizer suggests keeping the riser temperature at the upper limit ofthe riser

temperature while adjusting feed rate to keep the wet gas flow stay at the upper limit of

the wet gas compressor flow. This is different from the Summer Mode, in which the FCC

throughput reaches the upper limit while adjusting the riser temperature to keep the wet

gas flow stay at the upper limit ofthe wet gas compressor flow.

8.2.4 Optimal Solution Analysis

The optimization problem at hand is a complexed problem with 32 decision

variables and 70 constraints. The optimizer can not find a feasible solution each time. It

was estimated that the optimizer failed to give a feasible solution in about 40 percent of

all trials. Even the a feasible solution was found, it is not necessarily the optimal solution.

The mean and standard deviation of feasible solutions from random starting points are

listed in Table 8.13.

Table 8.13 The Mean and Variance ofthe Feasible Solutions.

Operation Mode No. of feasible Tme

solutions Solution

Mean Standard

Deviation

Summer Mode 100

Winter Mode 100

-1.24169'

-0.73421'

-1.16709

-0.69699

0.053868

0.035588

1: The the objective function used in the optimization is equal to dividing the revenue per day by-10'

187

Page 203: 5-95 gap chart

The magnitude ofthe standard deviation indicates that the solutions found by the

optimizer scatter around the real optimum in a fairly large area. In this work, about 200

sets starting points were used as starting points for each operation mode and the best

solution was regarded as the tme solution.

In order to have some confidence in the optimal solutions found in previous

studies, a depth test was condcuted to find out the shape of objective fiinction surface

around the optimal solutions. Tests were done to the optimal solutions of both Summer

Mode and Winter Mode. Tests were carried out by choosing some random points around

the optimal solution and calculating the objective fiinction value of each random point. A

random point was selected by perturbing the values of decision variables using the

formula given below:

X, = x^p,j + (0.5 - rand) - factor - {bu, - bl,), (8.5)

where

Xi- the i" decision variable ofthe random point,

Xopt,i- the i"' decision variables ofthe optimal solution,

rand- random number between 0 and 1,

factor- the factor, 0.04.

bui- the upper limit of i" decision variable,

bli- the upper limit of i* decision variable.

Since only the nearby surface was studied here, the change of a decision variable

was limited in the interval of ±2% of normal operating range. Since the optimization

problem at hand is a constrained problem, it is obvious that some random points

generated by equation 8.5 will have constraint violations. These random points were

discarded in the study. The Euclidean distance between each random point and the

optimal solution and the relative change in objective function value were calculated. The

decision variables were nomalized in calculating euclidean distances. It is found in the

tests that there is no random point that has a better objective fimction value than the

188

Page 204: 5-95 gap chart

corresponding optimal solution. This indicated that the optimal solutions found in

previous studies are at least the local optimums. The average euclidean distances and

average relative changes of objective fiinction are listed in Table 8.14.

Table 8.14 Change ofthe Objective Function Value around Optimal Solutions.

Operation Mode Average Euclidean Average relative change

distance of objective function

Summer Mode 0.049355 -0.00086

Winter Model 0.049834 -0.00418

It can be seen that the change ofthe objective fiinction values around the optimal

solution of Winter Mode is one magnitude larger than that of Summer Mode. Considering

the stopping criteria of l.E-6 for the optimizer, the changes of objective fimction values

for both modes are significant. Hence, it is concluded that the surface around each

optimal solution has enough steepness which will not cause the optimizer to stop before it

reach the vinicity ofthe solution. Why the optimizer stops in the vinicity of a tme

optimum remains unknown and ftirther is needed on this issue.

8.2.5 Profitability Improvement

The profitability improvements ofthe two operation modes are shown in Table

8.15. The average profitability improvement is 4.1%. The total revenue increase is about

1.3 million dollars per year. The refinery considered in this work is a small refinery with

a capacity of 50,000 barrels cmde oil per day. Considering the fact that some big

refineries have a capacity of 400,000 barrels cmde oil per day, the absolute profitability

improvement can be more significant. This profitability improvement is also comparable

to the profit of implementing advanced process control, 3-5% (Ellis, 1998), and that of

implementing real-time optimization, 3-5% (Ellis, 1998).

For the refinery considered in this work, the throughput ofthe FCC unit has an

upper limit of 20,000 barrels per day determined by environment regulations. Hence, no

189

Page 205: 5-95 gap chart

Summer Mode

Winter Model

Average

+4.5%

+3.6%

+4.1%

benefit ofthe throughput increase, which is common to other optimization projects, can

be obtained from the refinery-wide optimization.

Table 8.15 Profitability Improvement ofthe Refinery-wide Optimization.

Operation Mode NLP vs. Base Case Incremental Revenue

"~~~~~ $973,000/six months

$468,000/six months

$72I,000/sixmonths

Annual Incremental Revenue $1,341,000/year

8.3 Single-Unit Optimization

Currently, on-line optimizer has been implemented in several units in the

refineries to push the operation to against the most profitable constraints. The units which

already have the on-line optimizer implemented include crude unit, FCC unit, reformer,

etc. However, the optimizer implemented on a single unit only considers the economics

of that single unit instead ofthe economics ofthe whole refinery. It is very difficult for a

single-unit optimizer to generate an optimum solution which is consistent with those

given by a refinery-wide optimizer. This is because it is difficult to obtain appropriate

prices for the intermediate streams ofthe single umt.

Cmde Unit is used as the example to show the difference between the solutions of

single-unit optimization and refinery-wide optimization. Cmde unit is the first major

processing unit in the refinery. It produces several side-draw products that are the feeds

to the downstream processing units in a refinery. Usually the side-draw products are not

final products sold in the market. The prices ofthe side-draw products from the cmde

unit ofthe Summer Mode are estimated from Linear Programming (LP) shadow prices

and market prices. LP shadow prices and market prices are two main resources for

estimating the prices of intermediate streams. For a side-draw product, market price is

used whenever it is available. Otherwise, LP shadow price from the LP reports is used.

The prices ofthe side-draw products are listed in Table 8.16.

190

Page 206: 5-95 gap chart

Table 8.16 Prices ofthe Side-draw Products from the Cmde Unit, Summer Mode.

Side-draw Product Price, $/bbl Source

Cmde

LSR

Heavy Naphtha

Jet Fuel

Diesel

Gas Oil

Residue

13.97

13.93

17.45

17.44

16.35

11.95

9.22

Market Price

Market Price

LP shadow price

Market Price of Jet-5

Market Price of No. 2 diesel fuel

LP shadow price

Market Price of No. 6 fuel oil

Using the same objective fimction form as that of refinery-wide optimization

shown in equation 8.2, optimization was conducted using the same cmde unit model

which has been used in the refinery-wide model. The subset of decision variables and

constraints ofthe refinery-wide optimization corresponding to the cmde unit were also

used for the single-unit optimization ofthe cmde unit. The values of decision variables in

the optimum solution ofthe single-unit optimization ofthe cmde unit are shown in Table

8.17.

It can be observed from Table 8.17 that the optimization solution ofthe single-

unit optimization ofthe cmde unit always reaches the upper limit ofthe ASTM end point

of those high-value side-draw products such as heavy naphtha, jet ftiel and diesel. This

means the optimum solution simply maximizes the side-draw product with higher value.

It can be explained by the fact that side-draw product production has the dominating role

in revenue calculation. The utility usage in the cmde unit has secondary importance in the

optimization (Xu, 1998). In addition, the utility usage does have big change even when

the ASTM 95%o points vary. Hence, the maximizing the size-draw product with higher

value is the obvious solution ofthe single-unit optimization.

191

Page 207: 5-95 gap chart

Table 8.17 Prices ofthe Side-draw Products from the Cmde Unit, Summer Mode.

Side-draw Product Price of side-draw product, $/bbl

Optimum value of Active Limit single-unit optimization

Feed to crude unit

AT fiimace outlet temperature,

op

Light naphtha ASTM 95%

point, °F

Heavy naphtha ASTM 95%

point, °F

Jet Fuel ASTM 95% point, °F

Diesel ASTM 95%) point, °F

Heavy Vacuum Gas Oil TBP

end point, °F

13.97

13.93

17.45

17.44

16.35

11.95

50,000.000

650.000

230.000

385.000

525.000

670.000

1,080.000

Upper Limit

Lower Limit

Lower Limit

Upper Limit

Upper Limit

Upper Limit

Upper Limit

In order to make a fair comparison between single-unit optimization and refinery-

wide optimization, the optimization is conducted for the rest ofthe refinery while fixing

the decision variables ofthe cmde unit at the optimum value given by the single-unit

optimization. In other words, the subsets of decision variables and constraints

corresponding to the cmde unit are not included in the decision variable set and constraint

set of this optimization study. The solution of this optimization is compared to that of

refinery-wide optimization shown in Table 8.18 and Figure 8.4.

It can be observed from Table 8.18 and Figure 8.4 that the single-unit

optimization produces more diesel and less gasoline than the refinery-wide optimization.

This product slate change makes the revenue ofthe solution ofthe single-unit

optimization about 1.6% less than the revenue ofthe refinery-wide optimization. This

case shows that single-unit optimization may result in suboptimal solution for the point of

view of refinery-wide economics. Single-unit optimizer normally will maximize the

192

Page 208: 5-95 gap chart

volume of the product with highest estimated price, which may not be the optimum for

the plant-wide operation.

In order to make the single-unit optimization generates solution consistent with

refinery-wide economics, appropriate prices for the intermediate streams are necessary.

Sometimes, external constraints from refinery-wide economics are also needed (Jones,

1999). This also shows that refinery-wide optimization is superior to single-unit

optimization in the sense refinery-wide optimization eliminates the need to evaluate the

prices of intermediate streams. The refinery-wide optimization only uses market prices.

In order to realize the full potential ofthe refinery. It is beneficial to carry out refinery-

wide optimization.

Table 8.18 Comparison of Single-unit Optimization with Refinery-wide Optimization, Summer Mode.

Purchase, bbl/day

Cmde

Toluene

Ethanol

Isobutane

Sale, bbl/day

light naphtha

Butane

SNL spot

NL spot

Sub.Oct. spot

SNL rack

NL rack

Sub. Oct. rack

Propane

Refinery-wide Opt.

50000

200

370

1314

0

976

72

3009

0

3300

16000

3700

306

Single-unit Opt.

50000

200

371

901

0

943

151

1564

12

3300

16000

3700

341

Relative Deviation, %

0.00

0.00

0.33

-31.44

0.00

-3.40

109.04

-48.02

0.00

0.00

0.00

0.00

11.48

193

Page 209: 5-95 gap chart

Table 8.18 Continued.

C3 product

jet fiiel A

Jet ftiel 5

low sulfur diesel

No. 2 diesel

No.6 fiiel oil

Summary

Purchase, $/day

Sale, $/day

Utility, $/day

Revenue, $/day

Refinery-wide Opt.

1192

800

5000

10539

3332

2464

736338.6

884888.93

23467.000

125083.33

Single-unit Opt.

976

800

5000

11409

3607

2452

731165.183

876975.827

22740.000

123070.644

Relative Deviation, %

-18.13

0.00

0.00

8.25

8.24

-0.47

-0.70

-0.89

-3.10

-1.61

194

Page 210: 5-95 gap chart

(U

o

ion

ca N 'B

O H

U • 3

P <u bO 3

m D

C ' 0

zati

B O H

0 -O

1

b u 3

efi

&i

vo 0 Z

(N

0

z

0

13 C/1 (1> •a

o o -g & crt

O

a. c«

z o o. on

- I

z 3 ca 3

X

ca * - X -^ : c .2? -o. — ca

3

13

I-I u 3 m

3 D u

s 3 o N

O H

O

is •B y (/5 3 o S cn c/:i

^ § •4—> C O

O N

II e:o o 3 o CO O H

s o

u 00

I-I 3 CiO

o o o fN

o o o o

o o o 00

o o o so

o o o

o o o <N

o/iaa 195

Page 211: 5-95 gap chart

CHAPTER 9

CONCLUSIONS AND

RECOMMENDATIONS

The two main objectives of this work were:

1. Develop an overall nonlinear refinery-wide model for a ftiel-oriented refinery that

represents the refinery operation.

2. Carry out optimization studies to the overall refinery-wide model to find optimal

operating conditions that are inline with the actual practice.

This chapter is essentially a final commentary on how far successfiil we were in

meeting these objectives, and what is left for ftiture work.

9.1 Conclusions

Among the two major steps listed above, the first step required the maximum

amount of investment of time and effort. A first-principle, nonlinear steady-state refinery-

wide model was developed by building individual models for all processes in a fuel-

oriented refinery and integrating them into a refinery-wide model.

A simplified model was developed for the cmde unit. The atmospheric tower and

the vacuum tower in the cmde unit were modeled based on material balance, energy

balance and empirical correlations. The main function ofthe cmde unit model is to

predict the yields and properties ofthe side-draw products given the feed information and

cut points. Predictions ofthe simplified model agree with a rigorous ChemCad cmde unit

model on the volumetric flow rates ofthe side-draw products and the gains between

volumetric flow rates and cut points. It is concluded that the gasoline blending model

represents the actual process reasonably well.

A complete set of gasoline properties with eleven properties is calculated in the

gasoline blending model. In the octane blending model, interaction coefficient method

using chemical compositions ofthe blending stocks are applied. The interaction

196

Page 212: 5-95 gap chart

coefficients in the octane blending model were regressed from the blending data ofthe

refinery considered in the work. The Chevron's VPBI method is used in the RVP

blending model. Chevron's VPBI method is a proven method with reasonable accuracy.

The vaporization properties are calculated from the ASTM curves ofthe blending stocks

assuming ideal mixing. Other properties of gasoline are also calculated by summing the

corresponding properties ofthe blending stocks. Basic on the proven methods used in the

gasoline blending, it is concluded that the gasoline blending model represents the actual

process reasonably well.

The modeling ofthe FCC unit and reformer unit uses the existing models. Minor

modifications to both models to accommodate the specific features of the processes in the

concerned refinery. The kinetic parameters are used as the adjustable parameters to

benchmark the models against indusfrial data. We could not find data points spread over

the operating region since both processes are operated close to steady state. The FCC

model was benchmarked at a base case operating point. Since the FCC model had already

been proven to predict reasonably well in the full operating region in earlier studies

(Ellis, 1996; Taskar, 1996) and only minor changes were made in this work. It is

concluded that the FCC model can represent the FCC unit of this particular refinery

reasonably well in the full operating range. The reformer model was benchmarked at two

base case operating points corresponding to the low severity operation and the high

severity operation. By benchmarking the reformer model against two operating

conditions in different operating regions, it is expected that the reformer model should

give reasonable predictions in the fliU operating range ofthe unit.

The alkylation unit model was developed based on stoichiometric relations and

data from linear programming (LP) reports. Simplified models are developed for other

units in the refinery including rerun unit, ROSE unit, hydrotreater, gas plant, and diesel

hydrotreater, based on the data from the LP reports. The data from the LP reports include

the data of both operation modes. Summer Mode and Winter Mode. These units do not

change much in each operation mode. Hence the models are considered adquate by using

only normal operating data.

197

Page 213: 5-95 gap chart

TOTAL method and n-d-M method are used to calculate paraffins, naphthenes,

and aromatics (PNA) data ofthe FCC feed. These methods have been proven the most

accurate methods available in the literature. The feed characterization ofthe reformer unit

is based on the cmde assays, cut point ofthe reformer feed, and typical feed

compositions. The prediction ofthe PNA information ofthe reformer feed should be

more accurate than the prediction from empirical correlations. The cmde assays should

be used whenever possible. There may be some model mismatch on the composition of

individual chemical components in the reformer feed. The mismatch is deemed to have

an insignificant effect on refinery-wide optimization because volumetric fiow rates

instead of stream compositions have the utmost importance in refinery-wide

optimization.

The predictions from the overall refinery-wide model are close to the data in the

LP reports for both operation modes. Because the refinery operation follows the LP

reports, the model predictions are also close to the plant data. It is concluded that the

refinery-wide model represents the fuel-oriented refinery reasonable well.

The computational load is a big concern in refinery-wide optimization. The size

ofthe refinery-wide model is huge, with more than 22,000 lines of Fortran code, among

which 25%) are comment lines. Each objective function evaluation or each derivative

evaluation requires the solving ofthe entire model. The computational efficiency is

cmcial in obtaining a solution in reasonable time.

With the computational issue in mind, the refinery-wide model is coded in a

straight-through form to ease computational load. This simplification enables the model

to be solved in the range of 3-5 seconds on a 400 MHz Pentium II machine.

The optimization study of Summer Mode showed a profit improvement of 4.5%

over the normal operating conditions. This is realized mainly by producing more gasoline

and less diesel, and blending all light naphtha into gasoline. In addition, the FCC feed

reaches the upper limit while the wet compressor flow upper limit is active.

The optimization study of Winter Mode showed a profit improvement of 3.6%

over the normal operating conditions. Contrary to the Summer Mode, the profit

198

Page 214: 5-95 gap chart

improvement is realized mainly by producing more diesel and less gasoline. In the

optimum solution of Winter Mode, all light naphtha is blended into gasoline. Although

the total amount of gasoline decreases, the amount of Super Unleaded Gasoline increased

in the optimum solution. The FCC unit is operated at the highest severity while the wet

compressor flow limit is active.

The most important characteristic of an optimization study is to that the optimum

solutions make sense. By discussing with engineers from the refinery considered in this

work, it is found that these results are consistent with the operation strategy ofthe

refinery for both operation modes. One example is that the FCC unit in the refinery is

usually operated against the wet gas compressor constraint. We concluded that the

optimizer is able to find the optimum operating conditions that make the refinery

operation more profitable. The average profit improvement is 4.1%, which is equal to

about 1.4 million dollars per year for the particular refinery. This revenue improvement is

consistent with 3-5%) expected profit improvement (Hendon, 2000). It can be argued that

real profit improvement may not be exactly the same number showed here because the

model is not perfect. Nevertheless, this work proves that refinery-wide optimization can

generate reasonable optimum solutions that agree with the industrial practice and that is

in a quantitative form with plant-wide consistence instead of vague feelings about the

profit increase.

The single-unit optimization on the cmde unit shows that the solution of a single-

unit optimization may not be consistent with the optimum operation strategy for the

entire refinery. The case study shows that single-unit optimization on the cmde unit

resuhs in suboptimal operating conditions with 1.6% profit loss compared to the refinery-

wide optimization. The conclusion is that refinery-wide optimization is necessary in order

to realize the maximum profit from the plant-wide perspective. The success of a single-

unit optimization depends on the appropriate prices ofthe intermediate streams and

appropriate extemal constraints.

In conclusion, we would like to say that we completed the two major steps

outlined at the beginning of this chapter.

199

Page 215: 5-95 gap chart

9.2 Recommendations

Considering that current work is the first attempt to study the refinery-wide

optimization, following recommendations are made:

1. The effectiveness of process optimization is highly dependent on the quality ofthe

models used to simulate the process. Considering the complexity of refining

processes, to build detailed models for all the units in a refinery is out the scope of

this work. The cmde unit model can be improved by using strict tray-to-tray models

for the atmospheric tower and vacuum tower. The preheat train, preflash tower and

heat exchanger can also be modeled to develop a complete cmde unit model.

Although this kind of cmde unit model has not been seen in academia, this has been

accomplished in industry using commercial simulation software package.

2. The FCC ten-lump yield model should be expanded in order to predict the FCC

gasoline composition. The FCC gasoline is regarded as one lump in the ten-lump

yield model. Hence, it is impossible to calculate the PNA information ofthe FCC

gasoline, which is required by gasoline blending model.

3. A rigorous model needs to be developed for the main fractionator ofthe FCC unit in

order to have a complete FCC model. The cut points ofthe side-draw products ofthe

main fractionator are important decision variables, which are often adjusted by

engineers to change the refinery operation. The flooding constraint in the main

fractionator can be active sometimes during the operation. After the rigorous model

ofthe main fractionator is developed, the cut points may be included in the decision

variable set and the flooding constraint may be included in the constraint set.

4. Only simplified models were developed for units other than the four major units.

Hence, the next logical step in this work would be to develop detailed nonlinear

models for these units. The main benefit of doing this is that the constraints for these

units can be accurately modeled. These constraints may become active in refinery-

wide optimization. In the current model, only capacity constraints are considered in

200

Page 216: 5-95 gap chart

these units. Other constraints may be included in the optimization study after detailed

models of these units are available.

5. All the models in current refinery-wide model are in sequential-modular form.

Sequential-modular models have an inherent disadvantage of no accurate information

ofthe derivatives due to nested convergence loops. The noisy derivatives cause the

optimizer "wandering" during the search (AspenTech, 1998). It has been already

observed that the SQP algorithm based NPSOL software package (Gill et al., 1986)

often quitted searching without finding a solution in the present work. It is suspected

that the difficulty is caused by inaccurate derivatives. Open form approach has

become the standard in industry. One ofthe major advantages of open form approach

is that accurate derivatives can be calculated since each process is modeled by

equations without nested convergence loops.

6. More detailed refinery-wide model using sequential modular approach may not be

computational efficient due to the iterations required in model solving because ofthe

existence of recycle streams. The open form approach can solve the entire refinery-

wide model altogether without iterations. In addition, computational improvement

techniques, like algorithm for sparse matrices, can be applied for large open equation

model.

7. Building a detailed refinery-wide model with open form may include as many as one

million equations (Hendon, 2000). It is recommended to build open form single-unit

model as the first attempt. Using existing models in the process model libraries

(PML) of commercial software packages, such as AspenPlus, may be the easiest way

to constmct an open form refinery-wide model.

8. It is beneficial to include the specifications of all products produced in a ftiel-oriented

refinery. Only the product specifications of gasoline are considered in this work. The

specifications of other products, such as jet fiiel, diesel, No. 6 fliel oil, are not

included. The constraints associated with these specifications, such as the freezing

point of jet fuel, and the flash point and the pour point of diesel, are indirectly

represented by cut points in the current model. Including these specifications in the

201

Page 217: 5-95 gap chart

model should increase the accuracy of estimating the violations of product

specifications.

9. Sulftir balance becomes important as the specifications of sulfur content in all

products become more stringent. The capacity ofthe sulfiir recovering unit may

become an active constraint in refinery operation. Hence, effort should be spent on

modeling the sulfur fiow and distribution in a refinery.

10. Selecting appropriate cmde type according to market change can be important to

improve the profitability of a refinery. Hence, it is beneficial to include cmde

selection in the decision variable set ofthe refinery-wide optimization.

11. The comparison between nonlinear refinery-wide optimization and linear

programming (LP) will be critical in convincing industrial people that nonlinear

refinery-wide optimization is a better tool than LP in plainning and scheduling. To

fiirther current work, effort should be spent on developing a linear model

corresponding to the nonlinear model developed in this work.

202

Page 218: 5-95 gap chart

BIBLIOGRAPHY

American Petroleum Institute. Measuring , Sampling, and Testing Cmde Oil. Bulletin 2500, API, New York City, New York, January 1955.

Arbel, A., Huang, Z, Rinard, I. H., and Shinnar R. Dynamic and Control of Fluidized Catalytic Cracker. I. Modeling ofthe Current Generation of FCC s. Ind. Eng. Chem. Res., Vol. 34, No. 4, 1995, 1228-1243.

Aspen Technology Inc. Real-Time Optimization System Training Course Notes, Houston, TX, 1998.

ASTM. Standard Specification for Automotive Spark-ignition Engine Fuel. ASTM Standard D4814-96, 1997.

ASTM. Carbon Distribution and Stmctural Group Analysis of Petroleum Oils by the n-d-M Method. ASTM Standard D3238-85, 1985.

Beck, R. J. Resurgent Oil Demand, OPEC Cohesion Set Stage for Optimistic Outlook for Oil Industry at the Turn ofthe Centiiry. Oil & Gas Journal, 1999, 97(42), 49-62.

Beck, R. J. Earnings Plunges along with Oil Price in 1998. Oil & Gas Journal, 1999, 97(18), 56-58.

Bensabat, L. E. U.S. Fuels Mix to Change in the Next 2 Decades. Oil & Gas Journal, 1999, 97(28), 46-48,

Hess, F. H., C. D. Holland, R. McDaniel, and N. J. Tetiow. Solve More Distillation Problems: 7. Absorber-Type Pipestills. Hydroc. Proc, 56(5), 1977, 241-249.

Cutier, C. Personal Communication, Lubbock, TX, 1999.

Chung, C. and J. B. Riggs. Dynamic Simulation and Nonlinear-Model-Based Product Quality Control ofa Cmde Tower. AIChE Journal, 41(1), 1995, 122-133.

Dantzig, G. B. Linear programming and Extensions. Princeton University Press, Princeton, N J, 1963.

Daubert, T. E. Petroleum Fraction Distillation Interconversions. Hydroc. Proc, 73(9), 1994,75-78.

Dhulesia, H. New Correlations Predict FCC Feed Characterization Parameters. Oil & Gas Journal, Jan. 13, \9S6, 51-54.

203

Page 219: 5-95 gap chart

Edgar, T. F., and D. M. Himmelblau. Optimization of Chemical Processes. McGraw-Hill, Inc., New York City, NY, 1988.

Edmister, W. C. Improved Integral Technique for Petroleum Distillation Calculations. Ind andEngr. Chem., 47(9), 1955, 1685-1690.

Edmister, W. C , Lee B. I. Applied Hydrocarbon Thermodynamics. Gulf Publishing Company, Houston, TX, 1984.

Ellis, R. Supervisory Optimization ofa Fluidized Catalytic Cracking Unit. Master's thesis, Texas Tech University, Lubbock, TX, 1996.

Ellis, R. Personal Communication, Houston, TX, 1998.

Ellis R. C , X. Li, and J. B. Riggs. Modeling and Optimization ofa Model IV Fluidized Catalytic Cracking Unit. AIChE J., 44(9), 1998, 2068-2079.

Friedman, Y. Z. What's Wrong with Unit Closed Loop Optimization? Hydroc. Proc, 74(10), 1995, 107-116.

Hendon, S. Personal Communication, Lubbock, TX 2000.

Holland, C. D. Fundamentals of Multicomponent Distillation. McGraw-Hill, New York, 1981.

Gary, J. H., and G. E. Handwerk. Petroleum Refining: Technology and Economics. Marcel Dekker, Inc., New York City, NY, 1984.

Gill, P. E., W. Murray, and M. H. Wright. Practical Optimization, Academic Press, New York, 1981.

Gill, P. E., W. Murray, M. H. Wright and M. H. Wright. User's Guide for NPSOL (Version 4.0): A Fortran Package for Nonlinear Programming. Technical Report SOL 86-2, Department of Operations Research, Stanford University, CA, 1986.

Hartmann, J. C. M. Decision-making and Modeling in Petroleum Refining. Hydroc. Proc, 76(11), 1997,77-81.

Hess, F. H. Ph.D. Dissertation, Texas A&M University, College Station, TX, 1977.

Hess, F. H., C. D. Holland, R. McDaniel, and N. J. Tetiow. Solve More Distillation Problems: 7. Absorber-Type Pipestills. Hydroc. Proc, 56(5), 241.

204

Page 220: 5-95 gap chart

Holland, C. D. and Liapis A. I. Computer Methods for Solving Dynamic Separation Problems. McGraw-Hill Book Company, New York City, NY, 1983.

Huq, I. and M. Morari. Modifications to Model IV Fluid Catalytic Cracking Units to Improve Dynamic Performance. AIChE Journal, Vol. 41, No. 6, 1481-1499.

Jacob, S. M., B. Gross, S. E.Voltz, and V. W. Weekman Jr. A Lumping and Reaction Scheme for Catalytic Cracking. AIChE Journal, Vol. 22, No. 4, 701-713.

Jones, J. Personal Communication, Lubbock, TX 1999.

Lin, S. Petroleum Refining Engineering. Petroleum Industry Press, Beijing, P. R. China, 1988.

Lin, T.D. V. FCC Advanced Control and Optimization. Hydroc Proc, 56(5), 107, 1993.

Maxwell, J. B. Data Book on Hydrocarbons. D. Van Norstrand Company, Princeton, NJ, 1950.

Marlin, T. E. Process Control: Designing Processes and Control Systems for Dynamic Performance. McGraw-Hill, Inc., New York City, NY, 1995.

McFarlane, R. C , R. C. Reinemann, J. F. Bartere, and C. Georgakis. Dynamic Simulator for a Model IV Fluid Catalytic Cracking Unit. Comput. Chem. Eng., 17, 275-299, 1993.

McKetta, J. J. Petroleum Processing Handbook. Marcel Dekker, Inc., New York City, NY, 1992.

Mizoguchi, A., T. E. Marlin, and A. N. Hrymak. Operations Optimization and Control Design for a Petroleum Distillation Process. The Canadian Journal of Chemical Engineering. 73(6), 1995, 896-907.

Morris, W. E. Optimum Blending Gives Best Pool Octane. Oil & Gas Journal, Jan. 20 1986,63-66.

Morris, W. E. Gasoline Compositions in no-lead era. Oil & Gas Journal, Mar. 18 1985, 99-106.

Morris, W. E., W. E. Smith, and D. D. Snee. Interaction blending equations enhance reformulated gasoline profitability. Oil & Gas Journal, 92(3), 1994, 54-58.

Moran, M. J. and H. N. Shapiro. Fundamentals of Engineering Thermodynamics. John Wiley & Sons, Inc., New York City, NY, 1996.

205

Page 221: 5-95 gap chart

Nelson, W. L. Petroleum Refinery Engineering. McGraw-Hill, Inc., New York City, NY, 1958.

Palmer, F. H., A. M. Smith. The performance and specification of gasoline. E.G. Hancock (Ed.) Technology ofGasoline, Blackwell Scientific, London, 1985.

Pelham, R. and C. Pharris. Refinery Operations and Control: a Future Vision. Hydroc. Proc, 75(7), 1996, S9-94.

Pelham, R. Process optimization in the HPI. Hydroc Proc, July 1996, 89-94.

Riazi, M. R. and T. E. Daubert, Simplify Property Predictions. Hydroc Proc, 59(3), 1980,115-116.

Riggs, J. B. An Introduction to Numerical Methods for Chemical Engineers. Second Edition, Texas Tech University Press, Lubbock, TX, 1994.

Riggs, J. B. Chemical Process Control. Ferret Publishing, Lubbock, TX, 1999.

Sadeghbeigi, R. Fluid Catalytic Cracking Handbook. Gulf Publishing Company, Houston, TX, 1995.

Schoen, W. F. and A. V. Mrstik. Calculating gasoline blend octane ratings. Ind. and Engr. Chem., 47(9), 1955, 1740-1742.

Singh, A, J. F. Forbes, P. J. Vermeer, and S. S. Woo. Model-based real-time optimization of automotive gasoline blending operations. Journal of Process Control, 10, 2000, 43-58.

Symonds, G. H. Linear Programming Solves Gasoline Refining and Blending Problems. Ind andEngr Chem.,4%{3), 394-401, 1956.

Stewart, W. E. Predict RVP of Blends Accurately. Petroleum Refiner, 38(6), 1959, 231-234.

Stewart, W. E. Predict Octanes for Gasoline Blends. Petroleum Refiner, 38(12), 1959, 135-139.

Taskar, U. Modeling and Optimization ofa Catalytic Naphtha Reformer. Ph.D. dissertation, Texas Tech University, 1996.

Taskar, U. and J. B. Riggs. Modeling and Optimization ofa Semiregenerative Catalytic Naphtha Reformer, ^/C/z^Jowma/, 43(3), 1997, 740-753.

206

Page 222: 5-95 gap chart

Taylor, D. L., and W. C. Edmister. Solutions for Distillation Processes Treating ?etroleum Fractions. AIChE Journal, 17(6), 1971, I324-I329.

Turpin, L. E. Modeling of commercial reformers. Chemical Industries, 61, 1994, 437-480.

Twu, C. H. and J. E. Coon. Predict Octane Numbers Using a Generalized Interaction Method. Hydroc Proc, 75(2), 1996, 51-56.

Twoi, C. H. and J. E. Coon. Estimate Octane Numbers Using an Enhanced Method. Hydroc Proc, 76(3), 1997, 65-68.

Xu, J. S., personal communication, Houston, TX, 1998.

Watkins, R. N. Petroleum Refinery Distillation. Second Edition, Gulf Publishing Company, Book Division, Houston, TX, 1979.

Unzelman, G. H. Gasoline Volatility-Environmental Interactions With Blending and Processing. Fuel Technology & Management, May/June 1996, 39-44.

Vazquez-Esparragoza, J. J., Iglesia-Silva G. A., Hlavinka M., and Bullin, J. A. How to Estimate Rvp of Blends, Hydroc Proc, 71(8), 1992, 135-138.

Zahed, A. H., S. A. Mullah, and M. D. Bashir. Predict Octane Number for Gasoline Blends. Hydroc Proc, 72(5), 1993, 85-87.

207

Page 223: 5-95 gap chart

APPENDIX

CONSTANTS OF POLYNOMIAL EXPRESSON

Table A.l Constants of Polynomial Expression for the TBP Curve ofthe Mixed Cmde in Summer Mode, Calculating the Vol.% Given the Temperature in °F.

Order ofthe Polynomial Expression 5

Polynomial Constant Name Polynomial Constant Value

< 2

0.8957974413351622

-0.01372701021955436

0.0002952782429384016

-2.I49703441506334e-007

-3.854442121853092e-011

5.81465I604877458e-014

Table A.2 Constants of Polynomial Expression for the TBP Curve ofthe Mixed Cmde in Summer Mode, Calculating the Temperature in °F Given the Vol.%.

Order of the Polynomial Expression 5

Polynomial Constant Name Polynomial Constant Value

29.84428824723

25.62933368742233

-0.7922673091416073

0.01622221295752979

-0.0001531047698293264

5.656848165464662e-007

208

Page 224: 5-95 gap chart

Table A.3 Constants of Polynomial Expression for the TBP Curve ofthe Cmde in Winter Mode, Calculating the Vol.% Given the Temperature in °F.

Order of the Polynomial Expression 5

Polynomial Constant Name Polynomial Constant Value

2.193431671505095

-0.0401I70I235776854

0.0003645596718904187

-1.94657827079503e-007

-I.66907967834I765e-0I0

1.25674218569036Ie-013

Table A.4 Constants of Polynomial Expression for the TBP Curve ofthe Cmde in Winter Mode, Calculating the Temperature in °F Given the Vol.%.

Order ofthe Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

53.61544856365072

24.49778891041933

-0.8004076310844539

0.0170606968280822

-0.0001706104416978604

6.7218355I2472907e-007

209

Page 225: 5-95 gap chart

Table A.5 Constants of Polynomial Expression for the API Gravity Curve ofthe Mixed Crude in Summer Mode, Calculating the API Gravity Given the Vol.%.

Order ofthe Polynomial Expression 7

Polynomial Constant Name Polynomial Constant Value

< 2

119.7573532466777

-13.51427730842261

1.155003649306309

-0.05227710492067672

0.00129277077527945

-1.763855119596425e-005

1.24550423408025e-007

-3.55I0902548I0048e-0I0

Table A.6 Constants of Polynomial Expression for the API Gravity Curve ofthe Mixed Cmde in Winter Mode, Calculating the API Gravity Given the Vol.%.

Order ofthe Polynomial Expression 7

Polynomial Constant Name Polynomial Constant Value

115.3176682424892

-13.66362146845495

1.199745867594174

-0.05421826016288378

0.001324183745545327

-1.77568580195242e-005

1.2296309055042I6e-007

-3.43531951355496Ie-0I0

210

Page 226: 5-95 gap chart

Table A.7 Constants of Polynomial Expression for the Sulfiir Curve ofthe Mixed Cmde in Summer Mode, Calculating the Sulftir Given the Vol.%.'

Order of the Polynomial Expression 6

Polynomial Constant Name Polynomial Constant Value

-5.210428584234705e-005

7.080959174388113e-005

-3.525307004426104e-006

-I.0I7925910437589e-007

9.0797230957798I3e-009

-I.487796790868834e-010

7.430537669733543e-013

'The sulfur content is in wt.%).

Table A. 8 Constants of Polynomial Expression for the Sulfur Curve ofthe Mixed Cmde in Winter Mode, Calculating the Sulftir Given the Vol.%.'

Order ofthe Polynomial Expression 6

Polynomial Constant Name Polynomial Constant Value

~c 0.00030I4440208000I24

Cj -9.6I6598620532102e-005

c 1.523461 lI5826574e-005

c -8.795002067220337e-007

^ 2.239074678304I39e-008 ' ' 4

^ -2.49238838646I639e-010

r I.0I0096349284079e-012

The sulftu- content is in wt.%).

211

Page 227: 5-95 gap chart

Table A.9 Constants of Polynomial Expression for Converting ASTM End Point to TBP End Point. 1,2

Polynomial Constant Name Polynomial Constant Value

-89.99134827

2.74798583

-0.01068849

3.17354667D-5

-4.87231402D-8

3.72814245D-11

-1.11569087D-14

' The ranges that the polynomial expression applies to are: ASTM 95% point, 230-800°F; TBP end point, 250-850°F.

^The polynomial expression is regressed from the curves in Figure 2.15 of Watkins (1979).

Table A. 10 Constants of Polynomial Expression for Converting Gap (5-95) ASTM to Gap (0-100) TBP. 1,2

Polynomial Constant Polynomial Constant Polynomial Constant Polynomial Constant Name Value, hd-AGO Value, Id-hd Value, Ln-hn, hn-ld

'0

C4

^5

84.0867095

-1.16113063

-5.0455864D-3

1.2298931 lD-4

1.45938226D-7

-6.04374077D-8

116.80769309

-3.34042753

0.18403357

-8.27984365D-3

1.6959039 lD-4

-1.33260699D-6

171.90777015

-5.6022091

0.16631373

-3.16940048D-3

1.97107064D-5

0

' The polynomial expression correlates data when Gap (0-100) TBP is between 0°F and 100°F, and Gap (5-95 ) ASTM is between -20°F and 60°F.

^The polynomial expression is regressed from the curves in Figure 2.16 of Watkins (1979).

212

Page 228: 5-95 gap chart

Table A. 11 Constants of Polynomial Expression for Calculating the Molecular Weight of a Cmde Cut Given the Mean Average Boiling Point and API gravity, Characteristic Factor from 12.1 to 12.6.'"^

Order ofthe Polynomial Expression 5

Polynomial Constant Name Polynomial Constant Value

^ 2

-85.12057662010193

1.381740973796696

-0.003432246121519711

4.653977835999967e-006

-2.013575892828579e-009

2.29628573710579e-0I4

' The range of MeABP is from 200°F to 900°F.

^The polynomial expression is regressed from the curves in the figure on page 22 of

Maxwell (1950).

Table A. 12 Constants of Polynomial Expression for calculating the molecular weight of a Cmde Cut Given the Mean Average Boiling Point and API gravity. Characteristic Factor from 11.7 to 12.0.''^

Order ofthe Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

-64.81157898902893

1.187273759860545

-0.002771734283669503

3.45I7646I6357878e-006

-9.924137883005058e-0I0

-2.998295810238759e-0I3

' The range of MeABP is from 200°F to 900°F.

^The polynomial expression is regressed from the curves in the figure on page 22 of Maxwell (1950).

213

Page 229: 5-95 gap chart

Table A. 13 Constants of Polynomial Expression for Calculating the Molecular Weight of a Cmde Cut Given the Mean Average Boiling Point and API gravity. Characteristic Factor from 11.3 to 11.6. 1,2

Order of the Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

^2

C-,

-52.26850727200508

1.051935468334705

-0.002316809990588808

2.74I261329930467e-006

-5.501099888116645e-010

-3.779033623234492e-013

' The range of MeABP is from 200°F to 900°F.

^The polynomial expression is regressed from the curves in the figure on page 22 of Maxwell (1950).

Table A. 14 Constants of Polynomial Expression for Calculating the Enthalpy ofa Cmde Cut from Its API Gravity, Temperature and Phase. 1,2

Phase Liquid Liquid

API

Order ofthe Polynomial

Expression

Polynomial Constant Name

0

5

Polynomial Constant Value

58.86276904234546

0.350323675930667

3.154147293926712E-4

-2.422868286555691E-7

2.670881203784828E-10

-9.569643092979734E-14

10

5

Polynomial Constant Value

63.49363322873251

0.374694305181038

2.455940050722916E-4

1.791162849984485E-8

-6.447463588954254E-I2

-2.499495895424942E-15

214

Page 230: 5-95 gap chart

Table A. 14 Continued.

Phase Liquid Liquid

API

Order ofthe Polynomial

Expression

Polynomial Constant Name

20

5

Polynomial Constant Value

67.71119381915196

0.407442995888232

1.365238904540433E-4

3.40132834300566IE-7

-3.63881982759582E-10

1.372545297683309E-13

Phase

API

Order ofthe Polynomial

Expression

Polynomial Constant Name

^0

^1

C2

Ci

C^

n

Liquid

40

5

Polynomial Constant Vali

73.53287428134354

0.487630852423536

-4.193622I423I0005E-4

2.815101644570994E-6

-4.582843798569582E-9

2.585350426563243E-12

30

5

Polynomial Constant Value

70.58775514527224

0.445968287169308

-6.67553647062391lE-5

1.094693139080949E-6

-1.397221644590038E-9

6.189472668211555E-13

Liquid

50

6

7.852145754557569E+I

0.2332I2633516814E

3.992606531028287E-3

-2.54298013011578E-5

8.03349052350I33E-8

-1.174861450157626E-10

6.44705865814373E-14

215

Page 231: 5-95 gap chart

Table A. 14 Continued.

Phase Liquid Liquid

API

Order ofthe Polynomial

Expression

Polynomial Constant Name

60

6

70

3

Polynomial Constant Value Polynomial Constant Value

81.30727554178702 79.85688170857611

0.361148813299224

1.71I36428I39107E-3

-9.349315025275473E-6

3.02I385889742656E-8

-4.69397245I398978E-11

2.835893371433551E-14

0.460461386308936

2.079979360163353E-4

4.238046095624004E-7

Phase

API

Order ofthe Polynomial

Expression

Polynomial Constant Name

Co

C^

Ci

Ci

^4

^5

Ce

r

Liquid

80

7

Polynomial Constant Valu

84.04810924726189

0.290888623800129

8.617715997388586E-3

-I.55I777304769075E-4

1.423607972839136E-6

-6.656375318991881E-9

1.532741903504612E-11

-1.376303736593522E-I4

Vapor

10

5

185.5089059815509

0.3I06I1488926043

2.78I889429286366D-4

3.2I0699037192732D-8

-8.566150055727608D-1I

3.49229007131788D-14

216

Page 232: 5-95 gap chart

Table A. 14 Continued.

Phase

API

Order ofthe

Expression

Polynomial

Polynomial Constant Name

Co

c,

C2

Ci

c,

^5

Ce

Phase

API

Order ofthe Polynomial

Vapor

20

5

Polynomial Constant Value

206.0313633410842

0.263410338567155

5.549406963822889E-4

-6.I89165425302123E-7

6.05569I909972712E-10

-2.3688445303691 lE-13

-

Vapor

40

5

Vapor

30

6

Polynomial Constant Value

219.8464761304203

0.343124140861619

-1.409107793506337E-4

I.841500715915601E-6

-3.701660586125782E-9

3.545420054274547E-I2

-I.354470866I46257E-I5

Vapor

50

5

Expression

Polynomial Constant Name Polynomial Constant Value Polynomial Constant Value

^5

Ce

C-,

233.9651934063586

0.292765165784658

3.028I93795628908E-4

1.581314566512226E-7

-3.90264I490I37124E-10

2.14459I979966532E-13

244.699721654586

0.301180828185352

2.330962857755026E-4

3.511663374533569E-7

-6.03663435I300283E-I0

2.9780138812054I6E-13

217

Page 233: 5-95 gap chart

Table A. 14 Continued.

Phase Vapor Vapor

API

Order of the Expression

Polynomial Constant Name

60

5

Polynomial Constant Value

251.8660791841976

0.306490229043447

2.448736920257488E-4

1.660I99941688756E-7

-2.37550436I9027IE-I0

9.481283506091226E-14

70

5

Polynomial Constant Value

261.3058801981097

0.302175899050781

2.705655244454874E-4

1.169952889013004E-7

-2.028500223696939E-10

8.5602828I6441859E-14

Phase

API

Order ofthe Expression Polynomial Constant Name

Vapor

80

5

Polynomial Constant Value

275.80426423467

0.299026886717911

3.42965375921267E-4

-I.56378III4599649E-7

I.8I222870965312E-I0

-9.784376469162197E-14

Vapor

90

5

Polynomial Constant Value

299.5719446004368

0.30523I3I3810509

2.9157460I0856424E-4

2.7813798764I5357E-8

-8.450989602939215E-11

3.1I26I7977833836E-14

'The polynomial expressions correlate the data in the temperature range of 0°F to I200°F. The basis is liquid at -200°F. The unit of enthalpy is Btu/lb.

^The polynomial expression is regressed from the curves in Graph 1 in Appendix 2 of Watkins (1979).

218

Page 234: 5-95 gap chart

Table A. 15 Constants of Polynomial Expression for Calculating the Enthalpy ofa Cmde Cut in Saturated Vapor Phase from Its API Gravity and Temperature. '•

Phase Vapor Vapor

API

Order of the Expression

Polynomial Constant Name

40

6

Polynomial Constant Value

233.4883816781221

50

6

Polynomial Constant Value

243.9527180031873

c,

C2

Ci

CA

^5

Ce

Phase

API

Order ofthe Expression

Polynomial

r.

Constant Name

0.3285I8587I42753

-I.808624958812288E-4

2.632468991237147E-6

-6.258886072613157E-9

6.897744067547659E-12

-3.070086345068816E-15

Liquid

60

6

Polynomial ConstEuit Value

251.8582088107942

0.403622100740904

-1.852045328405438E-3

1.56524139338643E-5

-5.057826952398203E-8

7.465319567696318E-11

-4.157782127940892E-I4

Liquid

70

7

Polynomial Constant Value

262.0953662637621

0.338170381568489

-9.376740388233884E-4

1.340003429817216E-5

-6.I00737995340388E-8

1.2145I1902912453E-10

-8.96174754299945IE-14

0.217428358620964

2.309650226379745E-3

-2.73092370264294E-5

2.319865486377637E-7

-1.094177708507726E-9

2.487165117448216E-12

-2.138475785738538E-15

219

Page 235: 5-95 gap chart

Table A. 15 Continued.

Phase Vapor Vapor

"API 80 '- ~

Order of the Expression 7

Polynomial Constant Name Polynomial Constant Value Polynomial Constant Value

Co 276.035738839535

c, 0.279704957385547

c, 2.320989198778989E-3

C3 -6.030673040413603E-5

c 7.08380111014I508E-7

c, -4.129376077188773E-9 ^ 5

c 1.I56369803785671E-11

c -I.246418830627948E-14

'The polynomial expressions correlate the data in the temperature range of 0°F to 1200°F. The basis is liquid at -200°F. The unh of enthalpy is Btu/lb.

^The polynomial expression is regressed from the curves in Graph 1 in Appendix 2 of Watkins (1979).

220

Page 236: 5-95 gap chart

Table A. 16 Constants of Polynomial Expression for Calculating the Steam-free Delta T Given the Value ofthe Percent Stripout of Cmde Cuts. ''^

Constant Light Naphtha, Heavy Naphtha

Light Distillate Heavy Distillate, AGO, VGO

Order

0.04486391365208

1.876074961444829

0.070454820517625

6.651550673268503E-3

3.945026107317062E-4

1.150802438765197E-5

1.367411382319217E-7

0.047805972615606

1.472885452181799

0.047401526691829

-5.117099429298833E-3

3.396233130956716E-4

-1.063810533707965E-5

1.270384760859722E-7

-0.05627695162184

1.388574574821178

-0.0155528174116392

2.29546409696013 lE-3

-9.220926313879829E-5

1.397773963740523E-6

The polynomial expressions correlate the data in the percent stripout range of 10% to 30%. The unit of steam-free Delta T is °F.

^The polynomial expression is regressed from the curves in Figure 2.20 of Watkins (1979).

221

Page 237: 5-95 gap chart

Table A. 17 Constants of Polynomial Expression for Calculating Stripout from Steam Rate 1,2

Constant Light Naphtha, Heavy Naphtha

Light Distillate

c^ 26.14695580938133 ''0 c^ -142.7715174853802

c^ 983.3651920557022 ^ 2

^3 -4.343146883010864E3 c, 1.035283389568329E4 ''4

c^ -1.197023190975189E4

c 5.10I304268360138E3

Constant Heavy Distillate, AGO, VGO

c^ 42.21995899023023 ^0 c, -182.3823599107564 ' ' I

c^ 914.8338382840157

c^ -3.330140159606934E3

c 7.333609569549561E3 c -8.285482303619385E3 ''5 r 3.52004I457176209E3

32.79539154772647

-140.9899756833911

673.3640444278717

-2.076813027858734E3

3.610560988426209E3

-3.2039297351837I6E3

1.1I5012320518494E3

Reduced Cmde

36.73691331990995

-276.1504411064088

1.645269501239061E3

-6.296468217134476E3

1.382960I08709335E4

-1.533274863910675E4

6.3985597968I0I5E3

' The range of steam rate that the correlations apply is from 0 to 60 pounds of steam/barrel of stripped liquid. The unit of stripout is vol.%.

^The polynomial expression is regressed from the curves in Figure 2.13 Watkins (1979).

222

Page 238: 5-95 gap chart

Table A. 18 Constants of Polynomial Expression for Calculating the Steam-free DT Minus Actual DT(F) from the Temperature (F) Difference between Feed and Stripping Steam and Percent Stripout. ''

Percent 10% Stripout

20% 30%

Order 0.033722314836866

-1.065916488052709E-3

1.403707231880618E-6

-3.384139080733591E-8

-6.43010240533469E-12

1.4733666746051lE-13

4.334365325077462E-3

-7.77647455743441E-3

-7.249742002063996E-7

-2.896324521711872E-8

5.159958720330243E-12

-1.587679606245378E-15

0.025291736127647

-0.056599241882988

-3.732436294355723E-6

-9.642573628642823E-8

2.897515281416171E-11

-9.097404143844305E-14

'The polynomial expressions correlate the data in the range of-400°F to 400°F for the difference between feed temperature and stripping steam temperature. The unit of Steam-free DT is °F.

The polynomial expression is regressed from the curves in Figure 2.21 of Watkins (1979).

Table A. 19 Constants of Polynomial Expression for Calculating the Enthalpy of Steam from Temperature. 1,2

Order ofthe Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

4.070E+0

-0.616E-3

I.281E-6

-0.508E-9

0.0769E-I2

'The polynomial expressions correlate the data in the temperature range of 540°R to 1800°R. The unit of enthalpy is Btu/lb.

^The polynomial expression constants are from Table A-21E of Moran and Shapiro (1996).

223

Page 239: 5-95 gap chart

Table A.20 Constants of Polynomial Expression for Calculating the Enthalpy of Air from Temperature. 1,2

Order of the Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

-1.705163359642029

0.248415488749743

-I.90835I259771735D-5

1.393I9320524578D-8

5.89I37245932I254D-13

-1.166776I32260656D-15

'The polynomial expressions correlate the data in the temperature range of 500°R to 1320°R. The unit of enthalpy is Btu/lb.

^The polynomial expression is regressed from the data in Table A-22E of Moran and Shapiro (1996).

Table A.21 Constants of Polynomial Expression for Calculating the Enthalpy of Water from Temperature. 1,2

Order ofthe Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

-31.808

0.9984

'The polynomial expressions correlate the data in the temperature of I IO°F to 150°F. The unit of enthalpy is Btu/lb.

^The polynomial expression is regressed from the data in Table A-2E of Moran and Shapiro (1996).

224

Page 240: 5-95 gap chart

Table A.22 Constants of Polynomial Expression for Calculating the Slope of Flash Reference Line from Corresponding Distillation Reference Line.'

Order ofthe Polynomial Expression 5

Polynomial Constant Name Polynomial Constant Value

C2

Ci

c,

1.281878608097031 E-3

0.143415450581415

0.085872677521365

-2.583246I7I070641E-3

-2.19426I556383026E-4

1.534796214297529E-5

'The polynomial expression is regressed from the curve in the top fiugre on page 228 of Moran and Shapiro (1996).

Table A.23 Constants of Polynomial Expression for Calculating the Ratio of DT(flash)/DT(TBP) from Volumetric Percent Distillated. 1,2

Order ofthe Polynomial Expression

Polynomial Constant Name Polynomial Constant Value

0.206045789459949

0.098090319869243

-0.021268318498414

2.577717061399198E-3

-1.856466131933132E-4

7.70299445962408E-6

-I.685715965774048E-7

1.50049388647662E-9

'The volumetric percent distillated is in the range of 0% to 100%. The unit of temperature is°F.

^The polynomial expression is regressed from the curve in the top figure on page 229 of Maxwell (1950).

225

Page 241: 5-95 gap chart

Table A.24 Constants of Polynomial Expression for Calculating the Temperature Difference between the Distillation and Flash Reference Curves from the Slope of Distillation Reference Curve.''^

Case T50 (Distillation Reference T50 (Distillation Reference Curve) < 300 °F Curve) > 300 °F

Order ofthe Polynomial Expression

10 10

Polynomial Constant Name Polynomial Constant Value Polynomial Constant Value

Co

'10

0.046535948757082

-4.02163065969944

12.65744310617447

-12.82720935344696

5.004324972629547

-0.43010459802977

-0.194861501455307

0.060458100866526

-7.29041948216036E-3

4.I90444597043097E-4

-9.5I13I7557553412E-6

7.510650799842551

5.426912218332291

-19.98555105924606

26.42150974273682

-18.22189289331436

7.298733577132225

-1.746701300144196

0.253036742098629

-0.021768757607788

1.024311339278938E-3

-2.0320097I7822803E-5

'The slope is in the range of 1 to 12. The unit of temperature is °F.

^The polynomial expression is regressed from the curve in the middle figure on page 228 of Maxwell (1950).

226

Page 242: 5-95 gap chart

Table A.25 Constants of Polynomial Expression for Calculating the T50 of Flash Curve under Vacuum from the Pressure and T50 ofthe Flash Curve under Atmospheric Pressure. 1,2

Pressure 5 mmHg lOmmHg

Order ofthe Expression

Polynomial Constant Ntime

Pressure

Order ofthe Expression

Polynomial Constant 'Name

Polynomial Constant Value

-1422.40771484375

47.98374176025391

-0.601449653506279

0.00387019373010844

-1.40265651680238Ie-005

2.897231166087977e-008

-3.1809I3909009431e-01I

1.440629782690204e-0I4

25 mmHg

7

Polynomial Constant Value

-39.01131057739258

-1.205847382545471

0.03788168542087078

-0.0003145772279822268

1.342447035312944e-006

-3.07732755944201 le-009

3.6129467202283I3e-012

-1.705918532418649e-015

Polynomial Constant Value

993.2503967285156

-26.02331018447876

0.2790574356913567

-0.001563787373015657

5.065428638317826e-006

-9.480977780640387e-009

9.524453057838112e-0I2

-3.975758540364488e-015

50 mmHg

5

Polynomial Constant Value

-147.5445291846991

2.970037258812226

-0.0I730I7894876466

6.152585774188424e-005

-9.930441327229977e-008

6.058668425965486e-011

227

Page 243: 5-95 gap chart

Table A.25 Continued

Pressure 100 mmHg 200 mmHg

Order ofthe Expression

Polynomial Constant Name

"-1

c,

c.

Pressure

Order ofthe Expression

Polynomial Constant Name

Polynomial Constant Value

-39.28636805340648

1.059575435821898

-0.002797604639908968

l.I7I939102029285e-005

-I.94886755414I254e-008

209850868949959e-0Il

300 mmHg

5

Polynomial Constant Value

-46.21546807419509

1.752744925528532

-0.00828888484647905

3.486289179832625e-005

-6.448200663179693e-008

4.374364872I323I4e-0Il

Polynomial Constant Value

-38.44197991769761

1.383079235354671

-0.005222666145073163

2.I29300323439054e-005

-3.68I46I7920I2786e-008

2.341769013081753e-011

400 m m H g

5

Polynomial Constant Value

-29.64459969289601

1.454257622928708

-0.005063669466494503

2.092687345367494e-005

-3.797053613197043e-008

2.536417829122271e-01I

228

Page 244: 5-95 gap chart

Table A.25 Continued

Pressure 500 mmHg 600 mmHg

Order of the Expression 5 5

Polynomial Constant Name Polynomial Constant Value Polynomial Constant Value

"1

C-,

Ci

c,

-15.75690979277715

1.200692083148169

-0.002377156030320293

-19.58084477111697

1.597486192593351

-0.008274154270111467

9.557379709423941e-006 4.942406480523687e-005

-1.668142990407517e-008

1.0758158373411Ie-011

-I.46668I26477234e-007

2.125436487237804e-010

-1.201000766633333e-0I3

' The range of T50 that the correlations apply is from 50 to 500°C.

^The polynomial expression is regressed from the curves in Figure II-1-24 in Lin (1988).

229