You are on page 1of 23

Flow through packed beds

Group 11 Yuan Jia, Yan Li, David Hlavka April 29, 2009

Abstract

Our project studied the pressure drop across a packed bed and compared the data to the analytical solution governed by Erguns equation (Ergun, 1952). Different types of material and varying airflow rates were used to effect the range of our Reynolds number and subsequent pressure drop. We found that there is a significant relationship between pressure, frictional factor and Reynolds number with respect to flow rate and inlet pressure. We used the change in pressure to determine two constants (k1 and k2) that were then compared to Erguns equation. In the end we discovered that our experimental data was close to Erguns within 20%.

Introduction There are many engineering applications that utilize a packed bed. The forms more commonly implemented today are incorporated with systems involving adsorption of a solute, distillation, filtration and separation (e.g. water purification via waste removal). In an ion exchange or catalytic reactor, a single fluid (liquid or gas) flows through a bed of densely compacted granules (Geankoplis, 2003). Controlling the

desired pressure drop over the bed increases the diffusion efficiency of the reagents. In a filtration system, pressure drop is crucial because fluid flows through the filter medium by virtue of a pressure differential across the bed (Geankoplis, 2003). Studying the pressure drop across the bed is therefore essential to determine filtration efficiency and expected time span before excessive buildup of filtered material occurs. Packed beds can be made in various forms that are designed in accordance with their application. For our experiment, we used approximately 0.0085 m3 (0.3 ft3) of glass beads with a diameter 15 mm and copper-coated metal beads with a diameter of 4.5 mm. Other possible materials like sand were considered, but due to time constraints not tested. Our goal was to study the pressure drop of one fluid (air) flowing through a variety of packed materials. This started with approximately 0.0085 m3 (0.3 ft3) of glass beads at a diameter of 15 mm. After initial testing yielded no useable results, we determined the diameter of the beads to be too great to induce a readable pressure drop with our current gage. The only way to remedy this issue was to either install more accurate gages that would incur greater cost and take more time to order or use smaller sized beads. We decided to use 4.5 mm copper-coated metal beads, for these were cheaper than new gages and faster to obtain.

Once we knew how and what we were going to test, we needed to find a way to predict our results. Since pressure drop depends significantly on the Reynolds number, we needed an equation that described flow behavior in both the laminar and turbulent regions. We found that the Ergun equation (Ergun, 1952), which will be explained next, is the established theoretical and industry standard.

Formulation: In order to understand the Ergun equation (Ergun, 1952), one must first know a little about the formulation. (For detailed derivation see Appendix A) Ergun defined a

Reynolds number Rep that depends on the superficial velocity s (the volumetric flow rate divided by the column cross-section area), the equivalent diameter of the particle Fig. 1 Theoretical illustration of Reynolds number vs. friction factor

Dp, the dimensionless void fraction , and , the density of the fluid (Fig. 1).

(1)

Ergun also defined that the friction factor, fp depends on s, the pressure drop, p , and the length of the packed bed L:

2 1

(2)

There are two independent equations that deal with flow behavior that have different Reynolds numbers. When Rep < 10, the Blake Kozeny equation suggests that the friction factor depends mainly on the Reynolds number:

= 75

(1)2 3

(3)

When Rep > 1000, or in turbulent flows, the viscous force of the fluid is insignificant. This means the Burke-Plummer equation describes the friction factor as independent of the Reynolds number:

= 8

7 (1) 3

(4)

The Ergun equation (Ergun, 1952), which covers the entire range of flow rate, can then be obtained by assuming that the viscous losses and the kinetic losses are additive, therefore the result is:

= 75

(1)2 3

+8

7 (1) 3

(5)

After experiments with different packing material and different flow rate, Ergun determined the general form of the equation:

150

+ 1.75

(6)

Or in terms of pressure change - p:

150 (1)2 2 3

1.75 2 1 3

(7)

Where k1 = 150 and k2 = 1.75 are constants established via experimentation.


4

Apparatus Design and Calibration Before we could begin testing, we needed to create a device that would accurately measure our pressure drop. Prior to construction, we drew a sketch of what we wished to make, which included; flow paths, pressure gage and flow meter placement and bed location (Fig.1). With all our design decisions complete,

construction began.

We started by using an

apparatus that had been built by a previous group who were studying fluidization principles of fine Fig. 1 Apparatus Sketch

aggregate, predominately sand (see Appendix C). Their device consisted of a plenum for flow consistency and a column where the bed would be located. We needed to modify this design to fit our testing parameters. We began by attaching a PVC couple to the upper column, which we screwed a lid onto that had a 0.9525 cm (3/8 in) hole drilled into it and Fig. 2 Finished Apparatus piping attached. We did this so we would be

able to control how much air left the apparatus. After the upper portion was completed and sealed properly, we attached a flow meter and valve onto the base of the plenum to accurately monitor the flow going into the column. The last modification we made to the device was the addition of a pressure gage above our bed. Since there was already a port

made for a pressure gage below the bed, we merely mimicked this method and created a second port approximately 30.48 cm (12 in) above the base of the column. The finished apparatus can be seen in Fig. 2. After modification was completed, we began our testing procedure. We initially calibrated the system via mercury filled U-tube manometer, because we did not know the pressure gage sensitivity we would require. From our calculations we found our pressure difference to be approximately 689.5 Pa (0.1 psi), but we wanted to be sure our theoretical values were accurate before

implementing gages.

We connected the U-tube

manometer to the two ports above and below the packed bed, and measured the pressure change. Direct reading was about 689.5 Pa (0.1 psi) change over a bed of 24.13 cm (9.5 in) in height and 10.16 cm (4 in) in diameter. Once we had completed the U-tube Fig. 3 LabVIEW VI

manometer testing, we knew our pressure range and

acquired an appropriate gage from 0 to 41.4 kPa (0 to 6 psi). Next we calibrated the new pressure gage via LabVIEW and a tank draining operation. We used a similar set up and procedure as the tank draining experiment previously conducted in the semester. Our VI can be seen in Fig. 3. Finally in order to ensure that our seals were generally working to our desired specification, we did an initial calibration of the apparatus. To do this, we attached our pressure gage and filled the apparatus with air. We then let it sit for approximately ten minutes and observed any pressure changes. Since there were no large

observable deviations, we were safe to assume that the apparatus was properly sealed and functioning.

Experimental Procedure/Data

Before we could begin testing our pressure gradients, we had to calibrate our new pressure gage. To do this, we followed the same procedure as seen in the tank draining experiment done earlier in the semester. We filled a column with 500 mL of water until it reached a height of approximately 80 cm, taking a reading at every 500 mL addition. Once we filled the tank to our desired height, we allowed the tank to drain out the base and let the gage read the pressure change over time. This became our calibration data, which we then used to convert voltage to pressure over time (Fig. 4).

After the gage was calibrated, we then began testing our bed. We started with glass beads at a diameter of 15 mm that we poured to 24.13 cm (9.5 in). Our Fig. 4 Calibration Graph

LabVIEW VI was set for a time of 5 seconds

with an inlet pressure of 137895 Pa (20 psi). Once five seconds elapsed, we found no results. After we tried multiple pressure ranges ([137895 Pa (20 psi) 344738 Pa (50 psi)] at 68948 Pa (10 psi) increments), we determined that the current configuration was

inadequate to extrapolate any usable data. A smaller diameter bead was necessary to achieve the pressure changes.

We decided to move from 15 mm glass beads to 4.5 mm copper-coated metal beads. Again we initially set our VI to collect 5 seconds of data at 137895 Pa (20 psi) and moved up to 344738 Pa (50 psi) at 68948 Pa (10 psi) increments. Once we saw a pressure change, Fig. 5 Friction Factor vs. 1/Re

Fig. 6 Friction Factor vs. Re Number

Fig. 7 Pressure Drop vs. Velocity

we knew our configuration was appropriate for proper data collection. We moved the VI collection time up to 10 seconds and, starting at 137895 Pa (20 psi), moved up to 344738 Pa (50 psi) at 34474 Pa (5 psi) increments. This gave us enough data to see a significant

relationship between inlet velocity and pressure change and frictional factor and Reynolds number over the packed bed (Fig.5, 6 and 7).

When analyzing the first compiled data, we found a large error to be present (in excess of 1000%). Upon further inspection of our apparatus, we discovered that a flow meter intended for water was being used instead of a meter intended for air. We replaced the flow meter, but due to time constraints, could not get one that was able to read the ranges we were operating in. Because of this, we were forced to come up with a means of predicting the values of our flow rate at higher-pressure ranges. linear Since there was a present

relationship

between the inlet pressure and flow rate, we took the slope at two known values. We chose 0 Pa (0 psi) and 0 scfm and then 137895 Pa (20 psi) and 24 scfm, which gave us a constant of 1.2 to multiply by. With this, we were able to predict where

Fig. 8 Corrected Calibration

the flow rate would be at our

higher pressure (pressures exceeding 172369Pa (25 psi)). When we had the proper flow meter installed, we recalibrated our gage and found our conversion equation again (Fig. 8). Our RMS value for the calibration was 164, so our deviation was not too large. Since

we were operating at room temperature with standard humidity, we converted our flow rate from scfm to cfm via a one to one relationship using the standard conversion equation (8).
CFM = SCFM x 14.5 x460+68

(8)

Where 14.5 is standard atmospheric pressure, 460 + 68 is room temperature converted to Rankin, Ta = is ambient temperature and pa = Barometric Pressure. Once we started using the new flow meter for testing, we discovered that we only got usable results at pressure ranges above 206843 Pa (30 psi). Anything below that had too much noise interference to give us accurate results. Again we took

readings starting at 206843 Pa (30 psi) and moved to a maximum of 310264 Pa (45 psi). Anything above that was too great for Fig. 9 Friction Factor vs. Re Number

Fig. 10 Friction Factor vs. 1/Re

Fig. 11 Pressure Drop vs. Velocity


10

our system to handle, so we were unable to generate any higher pressures. After we gathered the data, we analyzed it and came to conclusions from there (Fig. 9, 10 and 11).

Discussion

When looking at the graphs, one can see that there is deviation between our data and the theoretical. Since we were working with such small changes in pressure (no greater than 689.5 Pa (0.1 psi), what appears to be large on the graph is not very large at all. Each error bar on the graph has a maximum value of 20%, so our error does not exceed 20% when comparing Reynolds numbers. Looking at the inverse Reynolds graph and taking the trend line gives us an equation from which we can get our k values (k1 = 650.61 and k2 = 1.1232). Comparing the experimental with the theoretical, we see that our values, while not exactly the same, are within an acceptable range that is around 20%.

The only area where there is a greater than 20% variance is in our inlet vs. velocity graph. Since we did not have the most accurate flow meter, this can explain the large deviation of our graph. Even though the deviation exists, our graph mimics the orientation of the theoretical curve, but is merely lower than expected. If we had enough time to acquire the appropriate flow meter, there is a greater possibility that the deviance would be much smaller.

Overall our test was a success, but there were many areas that contributed to error in our results. Most importantly was maintaining a full seal throughout the device. Since we had an apparatus that was previously used, the seals were not as strong and had the potential to leak or rupture due to degradation over time. During our initial

11

testing/calibration, the seal around the plenum ruptured due to a large internal pressure that rapidly accumulated from too much inlet pressure. Once the integrity of the seal was compromised, we had to reseal the entire apparatus to ensure no other seals burst during actual testing. As mentioned above, we attempted to check the seals to ensure accuracy in our results.

A second error was pipe leakage from the inlet.

This discovery was made when

we observed a large variance between our experimental and theoretical results. After further inspection, we determined leakage to be the cause. Instead of submerging the apparatus in water or using a soapy solution to observe bubble generation, we decided to use data variance as our check. We reduced the overall leakage by using Teflon tape on the connection threading and ensured that the fittings were tight. precautions, there were still minor leaks that could not be avoided. Even with these

A third significant source of error was found in our flow meter. As mentioned above, we were using the incorrect flow meter initially and this caused us to have a large error (approximately 1000%). Once we had the appropriate flow meter, it was too small for the flow ranges we were operating in. Since we did not have enough time to acquire a better meter, we came up with a linear correlation factor of 1.2 and used this to find our flow rates. We did not have enough time to obtain a better flow meter that could retrieve more accurate results so we had to assume our method was accurate enough. This error is most noticeable when correlating inlet velocity and pressure drop as seen in Fig. 11. A final source of error was the derivation of Erguns equation. Ergun derived his equation from experimental data that he acquired from repeated testing of various
12

materials in a packed bed fashion (Ergun, 1952). He did not account for all the variables, but made a general equation from his macro level observations. For one thing, he did not factor in surface roughness. Further experiments conducted by other groups determined that surface roughness does in fact alter the pressure, but not by a significant amount at the macro level (Kececioglu, 1994). Because of the derivation being from empirical data only, there is a level of error between our data and what Ergun derived. In the end we came to the conclusion that Erguns equation is generally true on the macro level, for our data. We did, however, find that there are a few inconsistencies present within the equation that in turn caused variances between our theoretical and actual findings. Further analysis could attempt to find just how inaccurate the equation can be.

Further Study

Due to time constraints, we were unable to complete all the testing proposed at the beginning of the semester. Because of this, there are many areas of further research that can be done to advance the understanding of flow over packed beds. Listed below are a few ideas that we thought would be interesting for further research.

As mentioned above, the first area of study could be trying different aggregates of varying size, shape and surface roughness to establish the inaccuracies present in Erguns equation. It would be interesting to see how a bed comprised of non-spherical particulate would behave compared to ours. Also studying finer aggregate (i.e. sand) would be another area of research. We attempted to study sand, but ran into difficulty establishing

13

an ideal packed bed and actual void fractions present. We ran out of time before we could come to any significant conclusions, but thought a vibratory rig would be appropriate for packing the material.

A second area that one can study is how a more viscous fluid behaves over a packed bed. These fluids can be as simple as water or as complex as heavy oils, but there would be some interesting results found when taking our testing procedure and applying it to various fluids. One could also try the varying fluids over different sizes of packed beds and aggregate, which would also yield unique results.

A final area of study would be to test various bed heights and widths at a consistent inlet velocity. Our findings seem to point to the fact that the taller the bed, the larger the pressure change and vice versa. We didnt, however, have the time to

empirically prove this true through testing, so any further testing in this area could have interesting conclusions.

14

References Carpinlioglu, M.O., Ozahi, E. A simplified correlation for fixed bed pressure drop Powder Technology 187 94-101 (2008). Ergun, S., Orning, A.A. "Fluid Flow Through Packed Columns", Chemical Engineering Progress. 48 89-94 (1952). Flow Through packed beds 20 February 2009 ftp://ftp.colorado.edu/cuboulder/courses/chen3200/Lectures/FlowAppl_3.pdf Geankoplis, Christie J. 2003 Transport Process and Unit Operations. 4th ed chapter 3 Prentice Hall: New Jersey. Guimard,P., McNermy,D., Saw.E., &Yang,A. 20 February 2009 http://rothfus.cheme.cmu.edu/tlab/pbeds/projects/t4_s04/t4_s04.pdf Kececioglu, I., Jiang, Y Flow Through Porous Media of Packed Spheres Saturated With Water Transactions of the ASME 116 164-170 (1994). Tully,A., Karasek,M., & Doraiswamy,S. 20 February 2009 http://www.me.rochester.edu/courses/ME241/11-Sand.pdf

15

Appendix A Excerpt from Transport Process and Unit Operations. 1 The void fraction is defined as:

( )

(1)

The specific surface of a particle av in m-1 is defined as

(2)

where Sp is the surface area of a particle in m2 and vp the volume of a particle in m3. For a spherical particle,

(3)

where Dp is diameter in m. For a packed bed of nonspherical particles, the effective particle diameter Dp is defined as

(4)

Since (1- ) is the volume fraction of particles in the bed

= 1 =

(5)

where is the ratio of total surface area in the bed to total volume of bed (void volume plus particle volume) in m-1. The average interstitial velocity in the bed is v m/s and it is related to the superficial velocity vs based on the cross section of the empty container by
1

Geankoplis, Christie J. 2003 Transport Process and Unit Operations. 4th ed chapter 3 Prentice Hall: New Jersey.

16

=
The hydraulic radius rH for the flow is defined as

(6)

= = = =

(7)

Combining equations (5) and (7),

= 6(1)
packed bed is as follows using equations (6) and (8).

(8)

Since the equivalent diameter F for a channel is = 4 , the Reynolds number for a

= 6(1)

= 6(1)

(9)

For packed beds Ergun defined the Reynolds number as above but without the 4/6 term:

= (1)
The Hagen-Poiseuille equation describes pressure drop in laminar flow

(10)

= (1 2 ) =

32 (2 1 ) 2

(11)

where p1 is the upstream pressure at point 1, N/m2; P2, pressure at point; v is average velocity in tube, m/s; D is inside diameter, m; and (L2-L1) is length of straight tube. Combining equation (6), (8), and (11) to give:
17

32 2

32 ( ) (4 )2

72 (1)2 3 2

(12)

The true is large because of the tortuous path and use of the hydraulic radius predicts too large a vs. which gives the Blake-Kozeny equation for laminar flow, void fractions less that 0.5, effective particle diameter Dp, and Reynolds number < 10.

150 (1)2 2 3

(13)

For turbulent flow in packed beds, the same procedure was used. First the pressure drop is defined as

= 4
Then combining equation (6), (8) and (14)

2 2

(14)

3 2 1 3

(15)

For highly turbulent flow the friction factor should approach a constant value. Also, it is assumed that all packed beds should have the same relative roughness. Experimental data indicated that 3f = 1.75. Hence, the final equation for turbulent flow for Reynolds number >1000, which is called the Burke-Plummer equation, becomes

1.75 2 1 3

(16)

Adding equation (13) for laminar flow and equation (16) for turbulent flow, Ergun proposed the following general equation for low, intermediate, and high Reynolds numbers which has been tested experimentally.

18

150 (1)2 2 3

1.75 2 1 3

(17)

Rewrite equation (17) in terms of dimensionless groups,

3 ( )2 1

= + 1.75

150

(18)

19

Appendix B Excerpt from A simplified correlation for fixed bed pressure drop.2 An open circuit blower type air flow test set up consisting of horizontal and vertical rigid pipes of inner diameter D of 103 mm (Fig.B1) was designed and constructed Air flow was generated by means of a fan in combination with an AC motor speed control unit (Siemens Micromaster Type MM 110). The packed bed location on the vertical pipe line was at Xv = 1050 mm from the 90 elbow. Packed beds were made up of uniformly sized spherical and non-spherical particles of irregular shape covering a range of (spherocity, where 1 is a perfect sphere) where 0.55 1.00 with the form of random loose packing. Crushed solid particles of natural zeolite, chick-pea and spherical particles of glass beads were used as packing materials. Packed beds of four different L providing a range of 0.24 L / D 1.46 were constructed. Bed limiters were made up of synthetic cloth with an open area ratio, = 0.92, wire diameter dw = 0.39 mm and mesh size M = 1.8 mm. The utilized packed bed codes, measured bed porosities and the range of the conducted test cases in terms of Rep are given in Table 2.

Fig. B1. A sketch of the experimental test set-up.

Carpinlioglu, M.O., Ozahi, E. 2008 A simplified correlation for fixed bed pressure drop Powder Technology 187(1) 94-101.

20

Table 1: Packed bed parameters and range of test cases


Packed bed code Z0640 Z0660 Z06150 C0925 C0940 C0960 C09150 G1625 G1640 G1660 Z1825 Z1840 Z1860 Z18150 Glass bead Dp = 16 mm 2675 Chickpea Dp = 9 mm 970 Packing material Zeolite Dp = 6 mm m (kg/m3) 640 L D / Dp Rep of test (mm) cases 40 17.16 0.82 0.37 1104, 1142, 1187 60 0.36 687, 950, 1075, 1125 150 0.36 675, 837, 912, 1009 25 11.44 0.55 0.44 1307, 2464, 2592, 2721 40 0.40 1200, 1850, 2180, 2300 60 0.40 1200, 1850, 2150, 2280 150 0.38 1596, 1761, 1838 25 6.43 1.00 0.44 2380, 4647, 4761, 5047 40 0.40 2168, 4177, 4408, 4462 60 0.39 2115 25 5.72 0.76 0.56 3954, 6818, 6981, 7772 40 0.48 3000, 5630, 5815, 6046 60 0.45 2792, 4887, 5192, 5541 150 0.40 2400

Zeolite Dp = 18 mm

640

21

Table 2: The measured mean exit velocity and pressure drop data for all test cases
Packed bed code Z0640 Z0640 Z0640 Z0660 Z0660 Z0660 Z0660 Z06150 Z06150 Z06150 Z06150 C0925 C0925 C0925 C0925 C0940 C0940 C0940 C0940 C0960 C0960 C0960 C0960 Re 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 U (m/s) 1.74 1.8 1.87 1.1 1.52 1.72 1.8 1.08 1.34 1.46 1.615 1.22 2.3 2.42 2.54 1.2 1.85 2.18 2.3 1.2 1.85 2.15 2.28 PBed (Pa) 430 447.9 532.4 289.6 546.7 706.8 862.5 689.2 1085.8 1005 1126 71.4 201.4 236.1 286.5 153.7 334.8 419.8 461.2 195.6 504.3 600.1 692.4 Packed bed code C09150 C09150 G1625 G1625 G1625 G1625 G1640 G1640 G1640 G1640 G1660 Z1825 Z1825 Z1825 Z1825 Z1840 Z1840 Z1840 Z1840 Z1860 Z1860 Z1860 Z1860 Re 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 15,000 25,000 30,000 33,000 U (m/s) 1.82 1.9 1.25 2.44 2.5 2.65 1.22 2.35 2.48 2.51 1.21 2.1 2.5 2.56 2.85 1.3 2.44 2.52 2.62 1.28 2.24 2.38 2.54 PBed (Pa) 982 1005 51.4 165.2 198.9 223.3 110.1 330.9 407.8 407.4 177.4 38.9 51.8 54.6 65.4 41.5 137.8 154.1 178.6 71.8 210.5 233.4 297.6

22

Appendix C Excerpt from Behavior during the Fluidization of Fine Sand. 3 Our first task was to build an apparatus that could hold our sand. The tank holding our sand was built of acrylic glass, since it was necessary to build the tank out of a clear substance in order to observe the height and behavior of the sand. Acrylic glass was readily available in the lab. Our tank was chosen with dimensions similar to those about which we had read. It is cylindrical with a 10 cm diameter and it is over a meter tall. This seems large, but the height is necessary to contain splashes from bubbling sand. Next, we needed to properly aerate the sand. Since our air comes from a compressor via small lines, we have a plenum designed to uniformly (roughly) push air into our tank, rather than shoot a jet of high velocity air directly through the center of the sand. This plenum is also made of cylindrical acrylic glass, and had to be partially machined, making it the second most expensive part of our experiment. We also had to create an effective seal between the plenum and the tank to keep the sand from falling into the plenum. The seal consists of a sheet of fiberglass supported by two perforated metal sheets and air tightened by two gaskets. Fiberglass is an ideal filter as it allows easy flow of air but still blocks our tiny sand particles. The tank is flanged, allowing it to be bolted into the seal and into the plenum.

Tully.A, Karasek.M, & Doraiswamy.S 20 February 2009 http://www.me.rochester.edu/courses/ME241/11-Sand.pdf

23

You might also like