Вы находитесь на странице: 1из 11

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

SIMULATION OF VISBREAKER FRACTIONATOR COLUMN STEP BY STEP PROCEDURES


Svetlin Vasilev, Dicho Stratiev1, Ivelina Shishkova
1

Lukoil Neftochim Bourgas - R&D Department, 8104 Bourgas, Bulgaria, e-mail: stratiev.dicho@neftochim.bg

KEY WORDS: process simulation ,visbreaker ABSTRACT


Simulating the performance of distillation columns whose feeds contain heavy oil components is the most challenging job. Poor feed characterization as well as inadequate column model techniques are the main reasons for inaccurate performance simulation. Computer models of systems processing wide boiling range hydrocarbon streams typically employ pseudo component representation of distillation fractions. In this method, commonly known bulk properties such as boiling point and gravity distributions are used in correlations to derive physical properties for petroleum fractions (pseudocomponents). These derived characteristics represent the properties of a mixture that is not, or can not be characterized by its individual chemical species. The heavy oil feed characterization is typically based on standard refinery laboratory distillation tests ASTM D-1160 or ASTM D-2887. These distillation tests end at a distillation temperature of 5400C and extrapolation of 5400C+ distillation data is applied. The proper extrapolation technique along with the proper correlations used for feed pseudocomponent characterization are critical for an adequate distillation column performance simulation. In this work an example of simulation of a visbreaker column performance is presented where all feedstock characterization and modeling techniques are addressed.

INTRODUCTION There are basically two main tasks a process engineer is faced with while creating a steady-state simulation model of a refinery fractionation column: 1. Specification of the feed requirements and proper thermodynamic models; 2. Creation of a flowsheet of adequately selected unit operations in order to create a mathematical model to closely match the real-life operation. The first task also involves the proper sample collection of the feed and products from the existing plant, and performing adequate laboratory analyses. It is well known among process designers that simulation of the performance of fractionators, whose feeds contain heavy hydrocarbon components is quite a challenging job. Poor feed characterization as well as inadequate column modeling techniques are the main reasons for the discrepancies between real-life operation and computer simulation, and also responsible for poor design of the plant [1]. It is known fact that computer models of systems employing hydrocarbons of wide boiling range are typically

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

represented with pseudocomponents, one pseudocomponent for a narrow boiling range fraction. Each pseudocomponents properties are predicted by its average boiling point and specific gravity specified at least. The heavy hydrocarbon fractions represent the main obstacle to fully describe all the products. This is based on the fact that the laboratory distillation test available (ASTM D-1160 and ASTM D-2887) are limited to distillation temperature of no higher than 540 oC. This requires the distillation data beyond this point to be extrapolated. In this article we will follow all the steps from the sample collection to the final simulation model that represents the real-life operation of the Visbreaker Fractionation Tower. We will also focus on how to deal with lack of laboratory data and how to overcome some of the short-comings, most commercial simulators have in the generation of pseudocomponents, extrapolating measured data and representing real-life phenomena that is not readily available in simulators unit operation models. SAMPLE COLLECTION AND LABORATORY ANALYSES The main obstacle in feed characterization is that it is impossible to directly take a good sample from the fractionators feed transfer line. The main reasons for this are: high temperature of the stream; presence of water steam in the feed; the flow regime in the transfer line is far from what we call ideal mixing; the stream is in vapour-liquid state (moreover the vapour and liquid are not in equilibrium) and in last place the cracking reactions which are still taking place in the transfer line. Obviously the only adequate way to describe the feed is by blending (mixing) all the product streams leaving the fractionator. This method is successful only if products flowrates are properly measured and maintained in steady condition for at least 24 hours. The following products, leaving the visbreaker fractionator can be identified: Hydrocarbon gas, visbreaker naphtha, process water, visbreaker gasoil (diesel) and visbreaker residue. Table I shows laboratory analyses, distillation method employed and measured product flow rates. As mentioned above it is difficult and even impossible to perform analysis of products with normal boiling points above 540 oC. Since the recovery of the material boiling up to 5400C in the visbreaker residue was about 40% it was necessary to extrapolate the data to get some additional data points. According to Kaes the most accurate approach to extrapolate distillation data is the use of probability distillation paper [2]. It was indeed found this to be the better alternative to using the routines for extrapolation in the process simulator. The end boiling point has been taken from the recommendations made by Kaes [2]. A probability distillation paper plot for Visbreaker residue of our case is represented in Figure 1.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Table I Physical and Chemical Properties of the Visbreaker main fractionation column products
Properties Density @15 0C kg/m3 Viscosity @80 0C, 0E Viscosity @ 100 0C, 0E Distillation, wt.% 8.3 12.2 15.0 21.4 29.1 34.7 50* 70* 90* 98* Visbreaker residue 992.9 33.6 11.6 TBP, 0C 360 400 420 455 490 517 565 640 755 850 D-2887 0 wt.% C IBP 101 5 153 10 164 20 180 30 196 40 212 50 228 60 244 70 262 80 281 90 308 95 331 FBP 395 D-86 0 vol.% C IBP 37 5 55 10 65 20 77 30 87 40 96 50 104 60 112 70 118 80 126 90 134 95 97 146 H2 CH4 CO2 C2H4 C2H6 C3H6 C3H8 C4H10 C5 C6 H2S vol.% 3.0 33.4 0.4 2.8 16.4 6.0 10.8 8.0 2.7 0.1 15.0 Visbreaker Diesel 827.2 Visbreaker Naphtha 714.0

Gas 1.302

Density @15 0C of the residue boiling above 5170C Flow rates, kg/h

1.0609 155 000 15 000 5 500 3 956

* Extrapolated from a probability distillation paper plot (see Figure 1) PSEUDOCOMPONENT BREAKDOWN AND GENERATION Two main issues should be considered when generating the pseudocomponents: 1) Blending of product streams to get the feed stream; 2) Proper prediction of pseudocomponents molecular weight. Some clarifications of the above-mentioned issues are given below: Pseudocomponent breakdown, generation and blending of dist. curves Let us first follow how pseudocomponents are generated in commercial simulators. The first step is to convert all the specified distillation curves in True Boiling Point (TBP) distillation curves. This is done with the use of empirical formulae or nomograms. If a single distillation curve is specified the computer simulator first divides the whole distillation range into the number of pseudocomponents specified (For more information of all the procedures, please consult your simulators user manual or Kaes [2]). Each pseudocomponent is given a normal boiling point an average of each interval and composition in volume or weight percent depending on the laboratory analysis type. Thus we defined one of the required properties of the pseudocomponent its normal boiling point. The other required property to calculate all the rest physical properties is the specific gravity. There are two available routines the simulator uses to calculate this property - the first and the most exact is by specifying an additional curve for each distillation curve specific gravity curve. However, this curve is not readily available from

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

laboratory analysis. So, we have chosen the second option, explained shortly below. In order to calculate each pseudocomponents specific gravity the computer simulator assumes equal Watson K-factor for each pseudocomponent and specific gravity is calculated from the Watson K-factor definition: (1) Watson K-factor = (NBP)1/3/sp.gravity, Where NBP is the normal boiing point of the pseudocomponent, in oR. Watson K-factor is determined iteratively so that the calculated overall specific gravity of the product matches the specified.

Figure 1.The probability distillation paper plot for the Visbreaker residue Blending of distillation curves is a procedure in which each pseudocomponent is given normal boiling point and sp.gravity specially fitted so that to correctly represent all the involved product distillation curves. In our case blending of distillation curves was avoided because it gave obsolete results. We adopted a method in which the Watson K-factor is calculated for each product distillation curve and assuming it is constant for all the pseudocomponents that belong to the same product, generated from this distillation curve. The pseudocomponents in our example are generated externally from the simulator and their properties are shown in Table II.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Table II Pseudocomponent breakdown and properties.


NBP 0 15 30 45 61 76 91 107 122 137 152 168 183 198 213 229 244 259 274 290 305 320 336 351 366 381 SG 0.6307 0.6420 0.6530 0.6636 0.6745 0.6845 0.6941 0.7042 0.7133 0.7697 0.7790 0.7887 0.7975 0.8062 0.8146 0.8235 0.8316 0.8395 0.8474 0.8555 0.8631 0.8705 0.8918 0.8991 0.9063 0.9133 MW 59.51 64.88 70.50 76.36 82.98 89.39 95.93 102.95 110.24 117.69 124.93 130.18 137.99 146.60 155.57 165.67 175.46 185.67 196.32 208.33 219.99 232.16 245.92 259.33 273.38 288.13 K-factor 12.51412 12.51412 12.51412 12.51412 12.51412 12.51412 12.51412 12.51412 12.51412 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.74121 11.562 11.562 11.562 11.562 NBP 397 412 428 444 461 477 493 510 526 543 559 576 592 608 625 641 661 684 707 730 753 776 799 822 845 868 SG 0.9207 0.9275 0.9347 0.9417 0.9491 0.9559 0.9627 0.9698 0.9763 0.9832 0.9896 0.9963 1.0025 1.0086 1.0151 1.0211 1.0285 1.0368 1.0451 1.0532 1.0612 1.0690 1.0768 1.0844 1.0920 1.0994 MW 304.83 321.18 339.23 352.86 374.68 396.40 419.58 446.18 472.94 503.87 535.26 571.89 609.51 651.00 700.43 752.57 828.19 934.18 1070.80 1196.00 1330.38 1478.85 1643.89 1827.35 2031.28 2257.97 K-factor 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562 11.562

Prediction of the pseudocomponents molecular weight Numerous methods exist for predicting the molecular weight of oil fractions. They can be divided into two groups graphical and analytical (empirical equations). The problem with most of these methods is that the derived correlations are generally made for predicting the molecular weight of straight-run oil fractions. When oil-fractions from conversion processes are specified discrepancies are observed from the actual molecular weight. The well know rule-of-thumb is: the higher is the average normal boiling point the greater is the deviation from the truth. In his work [3] Adriaan G. Goosens publishes a model applicable for straightrun fractions as well as products from conversion processes, where the standard deviation is only 2%. We adopted the correlation of Mr. Goosens in the calculation of the molecular weight of our pseudocomponents. The correlation stated in [3] is: MW = 0.01077(TBP) [1.52869+0.06486ln(TBP/(1078-TBP))]/d, (2)

Where d is the specific gravity @20 oC; TBP is the pseudocomponents or fractions average normal true boiling point in K.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Correct molecular weight is crucial for the calculation of the masstransfer and phase equilibrium in the fractionator since distillation curves are specified in volumetric or weight percent, but the phase equilibrium equations use mole percent. We have to point out that no single solution is available to a complex problem it can be seen that the heaviest last few components go beyond the correlation limit (above 804 oC the function is undeterminable as can be seen from Table 1). We extrapolated the function to get the molecular weight of these few pseudocomponents. The generated in this way pseudocomponents with their basic properties normal boiling point, specific gravity and molecular weight are presented in Table 2. For our simulation we used ChemCAD for Windows of Chemstations Inc, a very robust and convenient simulator. It has an option for adding user defined components, which we used to enter the pseudocomponents generated by the way mentioned above. SPECIFYING THERMODYNAMIC METHODS FOR THE SIMULATION A lot of models exist for the calculation of the vapor-liquid equilibrium constant K. Most of these equations represent a specially developed equation-of-state or an equation of corresponding states, some are purely empirical. In our case we used the Grayson-Streed equation of state because the visbreaker fractionator operates at low pressure. When selecting thermodynamic method the advice of the applied process simulators manual can be used. BUILDING THE FLOWSHEET AND PERFORMING THE SIMULATION Figure 2 shows the Visbreaker Fractionator configuration as it is operated at site. The computer flowsheet used for the simulation is represented in Figure 3. There are several phenomena that should be considered when building the flowsheet. S. Golden, et all describe in their work [1] the main issues with deep cut vacuum columns simulation. Basically the same issue applies here to greater or lesser extent. We will follow in the same order as presented in [1]. Transfer line flowsheet representation. The transfer line in the Visbreaker unit connects the Soaker Vessels with the fractionator feed nozzle. The flow regime in the line is of a mixed vapour-liquid phase. The pressure drop in the transfer line is a function of the velocity, physical properties of the oil and its physical layout. Things are more complicated by the cracking reactions still taking place in the transfer line. As a result of the transfer line pressure drop we know that the phase change is changing throughout the transfer line length. In most of the possible flow regimes the vapour and liquid phases in the transfer line are not uniformly mixed and liquid and vapour are not in equilibrium. The vapour entering the fractionators flash zone is highly superheated and additional volatilized by the steam available in the line. In our simulation model the transfer line is represented by a single theoretical stage.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Legend: C- Column; F Furnace; D Drum; S Soaker; AC Air Cooler: HE Heat exchanger Figure 2: Visbreaker Fractionator Column simplified P&ID

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

HC gas

3040 Nm3/h 1

VISBREAKER FRACTIONATOR SIMULATION MEASURED DATA SHOWN gas 5 16 10 11 T=37.9 C P=2.7 kg/cm2 37.2 m3/h reflux 13 15 NAPHTHA 12 5500 kg/h 35 PRODUCT 20 177 C 19 36 WASH ZONE 38 28 22 15000 kg/h PUMPAROUND 21 3040 Nm3/h water drain 34

Naphtha

5500 kg/h 2

LHT CGO

15000 kg/h 3

WATER 6 RESIDUUM 155000 kg/h 4 T=132 C 9

17 37

14

SH VAPOURS 7

379 C FEED 24 TRANSFER LINE NON-IDEAL MIXUP

26 27 ENTRAINMENT

FLASH ZONE T=358 C P=3.72 kg/cm2

33 25 30 T=231.8 C 31 WASH OIL 13 m3/h 18 m3/h

Measured tray temperatures: #17 - 200 C #23 - 284 C #27 - 332 C #28 - 339 C

8 23 29 residue RESIDUE 155 000 kg/h

Figure 3: Computer flowsheet - Simulation representation of Visbreaker Fractionator Configuration

Flash zone The vapour-liquid feed enters the column and experiences pressure drop at the column feed nozzle. After that vapour and liquid are separated and vapour rises up the column. Normally some sort of vapour-liquid separating device (vapour horn, multivane distributor or more complex devices provided by column internal vendors). In our case no special device is available but a simple splash baffle at the nozzle outlet. This results in high heavy oil entrainment with the rising vapour. This increases the requirement for wash oil to the wash zone of the column. The liquid entrainment actually exceeds the overflash. Wash zone At best the wash zone function is to fractionate the volatilized contaminants and de-entrain residue. It should also de-superheat the rising flash zone vapour. The main obstacle in simulating the wash zone is that one cannot specify liquid, leaving upper tray to go to a lower tray. This conflicts the column material balance. There is no problem to specify a pumparound where liquid from a lower tray is cooled and introduced to upper tray. To overcome this wash zone is represented externally from the column unit operation.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Remaining trays of the fractionator. The remaining part of the column does not constitute significant problem. Distillation unit operation is used with Inside-out rigorous distillation method used. The column is fed with the overflash vapour from the wash zone, has a top product, one side-drawoff and liquid bottom product, which enters the wash zone. DISCUSSION OF THE SIMULATION RESULTS In Table III a comparison between measured plant data and simulation results is made. A close match can be seen between the computer simulation and the real plant data. Now as we match the computer model to reality we can highlight more important design task: 1) Reduce the consumption of the wash-oil in wash zone by choosing a column internals vendor to supply a vapour-liquid distributing device. 2) Increase the diesel drawoff flowrate. The visbreaker diesel is valuable feedstock to the hydrodesulphurization plant. Now it is mixed back with the residue to reduce its viscosity so it can be used as fuel oil. A possible alternative can be the FCC unit heavy cycle oil some preliminary calculation had been made. Table III Comparison between simulated and real performance of the visbreaker main fractionator
Simulated Column 0 profile, C Top 17 tray 23 tray 28 tray Bottom temperature 144 206 253 341 358 132 200 284 339 358 Real measured

Product quantity, t/h Gas Naphtha Diesel Residue Product quality Specific gravity Naphtha Diesel Residue Distillation 0 ASTM D-86, vol.%/ C Naphtha IBP 5 10 20 30 40 50

4.3 5.0 14.3 156.2

4.0 5.5 15.0 155.0

0.704 0.847 0.989

0.714 0.8272 0.993

-18 32 52 75 91 104 113

37 55 65 77 87 96 104

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

10

60 70 80 90 95 97 100 Diesel IBP 5 10 20 30 40 50 60 70 80 90 95 97 100 Residue TBP at 1 atm IBP 5 10 20 30 40 50 60 70 80 90 95 97 100

Simulated 120 126 132 141 151 163

Real measured 112 118 126 134 146

120 191 198 209 220 235 253 282 319 332 349 371 442

101 153 164 180 196 212 228 244 262 281 308 331 395

61 327 356 440 490 533 557 587 622 734 769 808 881

323 379 448 497 538 565 612 650 694 751 796

REFERENCE 1. Golden S. W., Villalanti D. C., Martin G. R., Feed Characterization and deep cut vacuum columns: simulation and design. Impact of High Temperature Simulated Distillation, AIChE 1994 Spring National Meeting, 1994, Atlanta, Georgia, USA. 2. Kaes G. L., Modeling of Oil Refining Processes and Some Practical Aspects of Modeling Crude Oil Distillation, VMGSim Users Manual, www.virtualmaterials.com. 3.Goossens A. G., Prediction of Molecular Weight of Petroleum Fractions, Ind.Eng. Chem. Res., 35 (3), 985-988, 1996.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

11

4. Bridjanian H., Ghaedian M., Hashemi R., Mohammadbeigy Kh., Effective parameters of PNA reduction in visbreaker naphtha, Petroleum and Coa,l 47 (3), 1-5, 2005. 5. Goossens A. G. Prediction of the Hydrogen Content of Petroleum Fractions, Ind. Eng. Chem. Res., 36, 2500-2504, 1997.

Вам также может понравиться