You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/228695763

Downer reactor: From fundamental study to industrial application

Article  in  Powder Technology · March 2008


DOI: 10.1016/j.powtec.2008.01.022

CITATIONS READS

90 300

5 authors, including:

Yi Cheng Changning Wu
Tsinghua University Southern University of Science and Technology, Shenzhen, China
218 PUBLICATIONS   3,004 CITATIONS    38 PUBLICATIONS   504 CITATIONS   

SEE PROFILE SEE PROFILE

Jesse (Jingxu) Zhu


The University of Western Ontario / Western University
583 PUBLICATIONS   9,036 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Electrostatic powder spray View project

organic optoelectronic materials and devices View project

All content following this page was uploaded by Changning Wu on 05 May 2020.

The user has requested enhancement of the downloaded file.


Available online at www.sciencedirect.com

Powder Technology 183 (2008) 364 – 384


www.elsevier.com/locate/powtec

Downer reactor: From fundamental study to industrial application


Yi Cheng a,⁎, Changning Wu a , Jingxu Zhu b , Fei Wei a , Yong Jin a
a
Department of Chemical Engineering, Beijing Key Laboratory of Green Chemical Reaction Engineering and Technology,
Tsinghua University, Beijing 100084, PR China
b
Department of Chemical and Biochemical Engineering, University of Western Ontario, London, Canada ON N6A 5B9
Available online 8 February 2008

Abstract

Downer reactor, in which gas and solids move downward co-currently, has unique features such as the plug-flow reactor performance and
relatively uniform flow structure compared to other gas–solids fluidized bed reactors, e.g., bubbling bed, turbulent bed and riser. Downer is
therefore acknowledged as a novel multiphase flow reactor with great potential in high-severity operated processes, such as the high temperature,
ultra-short contact time reactions with the intermediates as the desired products. Typical process developments in industry have directed to (1) the
new-generation refinery process for cracking of heavier feedstock to gasoline and light olefins (e.g., propylene) as by-products; and (2) coal
pyrolysis in hydrogen plasma which opens up a direct means for producing acetylene, i.e., a new route to synthesize chemicals from a clean coal
utilization process. This paper is to give a comprehensive review on the development of fundamental researches on downer reactors as well as the
particular industrial demonstrations for the fluid catalytic cracking (FCC) of heavy oils and coal pyrolysis in thermal plasma.
© 2008 Elsevier B.V. All rights reserved.

Keywords: Downer reactor; Hydrodynamics; Modeling and simulation; Fluid catalytic cracking, Coal pyrolysis

1. Introduction inevitably benefit the reaction selectivity and yields (Reh [7];
Gartside [8]). The concept of gas–solids downer reactor was
Fluidization technique has been developed for more than accordingly proposed (Gross and Ramage [9]), and attracted
several decades (see review articles, e.g., Grace [1,2]; Bi and great attention from industry, which can be disclosed by the
Grace [3]; Lim et al. [4]; Zhu et al. [5]; Kwauk et al. [6]). In patents on downer reactors by almost all the major oil companies
conventional gas–solids fluidized bed operations, particles are (Murphy [10]; Gross and Ramage [9]; Gross [11]; Niccum and
suspended by upflowing gas streams against the force of gravity, Bunn [12]; Dewitz [13]; Muldowney [14]; Pontier et al. [15]).
such as in bubbling fluidized beds, turbulent fluidized beds, and Since gas and solids move downwards co-currently in a downer
risers, which results in many advantages to the processes, e.g., reactor, backmixing of phases is effectively reduced thanks to
enhanced mass and heat transfer and improved inter-phase the assistance of gravity force. This in turn creates a new flow
contact, but also leads to heterogeneous flow structure and regime in operating gas–solids fluidized systems. Downer
significant backmixing of phases. The overall performance of becomes a unique reactor approaching to a plug-flow operation
fluidized beds which approaches to a continuous stirred tank for fast reaction processes among the multiphase reactors.
reactor (CSTR) operation limits further improvement in some Considering the inherent nature of downer reactor, two types of
specific processes, for example, the processes with demand on potential applications become outstanding. One is directed to the
the selectivity of intermediate product(s). fast catalytic conversions of feedstock such as heavy oil and
To achieve highly selective catalytic conversions and to avoid other hydrocarbons, in which solid particles are catalysts and
unwanted by-products in the industries such as FCC, a reactor the reactions always involve a series of consecutive/parallel
allowing for very short contact time between phases and pre- competing ones. Meanwhile, the particles are also heat-carriers
cisely controlled gas and solid catalyst residence times would to heat the feedstock. By taking advantages of downer reactor, a
promising HSCC (high-severity catalytic cracking) process
⁎ Corresponding author. Fax: +86 10 62772051. was proposed to produce the lighter olefins by deep catalytic
E-mail address: yicheng@tsinghua.edu.cn (Y. Cheng). cracking directly from heavy oil or residues (Aitani et al. [16]).
0032-5910/$ - see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2008.01.022
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 365

Another type of applications is the pyrolysis process of solid separated from the catalysts at the bottom of the reactor and
materials, in which reactions take place on particles at high quenched mostly to terminate the further reactions. If necessary,
temperatures and intermediate gas products are desired so that the catalysts need to be regenerated and circulated to the top of the
rapidly heating and quenching the multiphase mixture become the reactor. A distinct feature of the downer operation is that gravity
dominant factors in influencing the reactor performance. Typical assists the downflow streams so that uniform flow structures and
examples are pyrolysis of biomass, coals, solid wastes and so on. narrow residence time distributions (RTDs) are achieved for both
The great potential in industry and the unique multiphase phases. The plug-flow reactor performance of downer makes it a
flow phenomena of downer reactors stimulated the fundamental candidate for accommodating very fast reactions with the
researches over the world in 1990s. Zhu et al. [5] gave an intermediate(s) as the desired product(s), where the uniform
excellent review on the early stage researches on downer short contact time is essential to prevent over-reaction and to
reactors. Thereafter, a large amount of researches has been ensure good selectivity. These short contact time processes
carried out extensively, trying to reveal any details of downer generally react in milliseconds to seconds. From a reaction
reactors related with the hydrodynamics, mixing, heat transfer, engineering viewpoint, such processes put strong demands on the
modeling and simulation, lab-scale hot model test and so on. solids feeding and gas–solids mixing at the downer inlet as well as
Some industrial demonstration plants have been built up to the fast separator at the outlet. Thus, the reactor performance of
invest the practical processes of downer reactors such as in FCC downer would be very sensitive to the structure design, although
processes towards improved products distribution and in coal the concept design shown in Fig. 1 looks rather simple. On the
pyrolysis for acetylene. other hand, multiphase flows in downers exhibit unique features,
This paper is based on the presentation record by Prof. Yong which also challenges the predictability of modeling and
Jin who received the 2006 Lectureship Award in Fluidization simulation on the gas–solids flows. Considering the diversity of
issued by the Particle Technology Forum (PTF) of AIChE and fast reaction systems, the downer design is still very much
sponsored by PSRI (Particulate Solids Research Inc.). Some dependent on the experience or empirism, which definitely
additional materials are provided for a better overview on this increases the risk when exploiting downers into practice.
subject. Considering the review article by Zhu et al. [5], the
following content will mostly focus on the recent efforts made 3. Practical issues on downer design
in developing both fundamental understanding and process
development of downer reactors. 3.1. Inlet configuration for rapid gas–solids contact

2. Conceptual design of a downer reactor As mentioned above, the ultra-short contact times between
phases in the whole downer (e.g., less than 1 s) impose stringent
As a start, the essential concept design of a downer reactor is demand on the solids feeding and rapid gas–solids mixing at the
illustrated in Fig. 1. Feed (gas) meets catalysts (solids) at the top of downer inlet. In other words, the solids must be rapidly dispersed
a vertical column and the mixture flows downwards under both in the gas flow, having intimate contact with gas. Otherwise, the
gravity and the momentum of the entering flows. Products are non-ideal distribution of highly concentrated gas–solids flows at
the initial stage would cause fast reaction rates at local positions,
directing to the worsen products distribution.
In principle, the inlet distributor of downer should serve
uniform distribution of phases, quick acceleration of solids, and
excellent control of gas–solids mixing to make best use of the
advantages of downer reactors. However, the major challenge to
promote the initial gas–solids mixing comes from the gravity
acting on the particles. Entering solids tend to drop immediately
downward and cannot reflux or circulate so as to enhance the
initial heat and mass transfer with the gas. Gas usually flows
along a less resistant path, i.e., following the region of low
solids density. While, gases can be easily distributed in com-
parison with solids. Therefore, the major concern has been put
on the design of solids distributors. Fig. 2 summarizes the
representative inlet geometries reported in the literature.
The simplest design may come from the earlier researches on
downer hydrodynamics conducted at Fluidization Lab of
Tsinghua University (FLOTU). As shown in Fig. 2 (a), gas–
solids flow is not separated at the riser top and enters into the
downer directly through a 90° sharp bend (Qi et al. [17]), or a
smooth connection. In this case, gas and solids keep high
turbulent intensity and initial momentum at the downer inlet. But,
Fig. 1. Conceptual design of a downer reactor. the solids may not be well distributed when entering the downer.
366 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

Fig. 2. Inlet configurations of downer reactors in the literature. (a) Qi et al. [17]; (b) Cao et al. [18]; (c) Lehner and Wirth [22]; (d) Briens et al. [23]; (e) Muldowney
[14]; (f) Jin et al. [24].
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 367

Fig. 3. Gas–solids fast separators. (a) Gartside [8], (b) Qi et al. [26], (c) Wei et al. [28].

As the first prototype, this design allows easy construction and have approximately the same velocity at the end of the annular
operation of a riser–downer system. Basic hydrodynamic gap. Primary air and the gas/solids flow leaving the annular gap
behaviors were studied under this downer design, though the are mixed at the end of both pipes. A diffuser connects the outer
structure itself is not applicable for industrial practice. pipe and the downer. With this setup, backmixing at the
To re-distribute gas and solids at the downer inlet, a fluidized entrance of the downer can be suppressed.
bed feeder was designed, see Fig. 2 (b). Solids are fluidized It is acknowledged that good gas–solids contact in the early
uniformly above the downer inlet and flow through several tubes stage of reactions is very crucial to downers because of the
into the downer. Gas is introduced through the ring-slots around comparable residence time in the entrance region to the whole.
the tubes or from the side-wall of the downer (e.g., Cao et al. However, the contact efficiency is not equivalent to the uniformity
[18]). In this way, independent operation of gas flow rate for the of spatial distribution of gas and solids. Briens et al. [23] designed
riser and the downer can be implemented; solids flow rate can be a mixing chamber, which equipped with eight jet nozzles supplied
adjusted flexibly by the (fluidized) bed height, flow rate of with constant air mass-flow rates from eight calibrated sonic
fluidization air and the number of distribution tubes. Further- nozzles. The eight jet nozzles through which the gas enters the
more, this type of distributors can be easily designed and scaled mixing chamber can be oriented independently. Obviously, the
up so as to be suitable for the industrial development. Similar configuration plays a significant role in the gas–solids mixing and
ideas were employed by Johnston et al. [19,20], Ma and Zhu [21] contact. Fig. 2 (d) shows how the nozzles can be angled in the
with more careful considerations on the distributor details, e.g., horizontal plane to induce a swirl and how the nozzles could be
where and how to arrange the gas and solids injection points. inclined to hit the solids jet near its top or near the bottom of the
Lehner and Wirth [22] hired a different type of inlet structure mixing chamber. This mixing chamber provides more options in
as shown in Fig. 2 (c). The main feature of the structure consists controlling the early contact between phases.
of two concentric pipes, which are located in the center of the Different from above designs, Muldowney [14] invented a
distributor. Gas is sucked through the inner tube for primary air. process and apparatus for short contact time fluidized catalytic
The solids are fed to a fluidized bed with a screw feeder. The top cracking of heavy oil feed using a reactor with an upflow cat:oil
of the outer tube acts as a weir for the fluidized bed. The vaporizer and a downflow reactor, as shown in Fig. 2 (e). The
overflowing solids are fed into the annular gap, which surrounds basic ideas are (1) catalyst/oil mixing is more effective in upflow
the primary air tube. Secondary air can be added to the and (2) catalyst/oil reactions have advantages in downflow. This
distributor in order to accelerate the solids in the annular gap. It reactor design permits short contact time cracking of heavy feeds
is assumed that gas (secondary air and fluidizing air) and solids with high activity zeolite catalysts, and achieves both vigorous
368 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

catalyst/oil mixing and predominantly catalytic cracking without Two fast separators for gas and solids were presented in the
the influence of overspent catalyst anywhere in the reactor. early literature. One is a one-quarter turn cyclone gas–solids
Likewise, Jin et al. [24] proposed a novel reactor concept by separator (Gartside and Woebecke [25]; Gartside [8]). A sepa-
utilizing the advantages of both riser and downer reactors to ration time of 30 ms and an efficiency of 98% were claimed.
reduce the sensitivity of downer reactors to the inlet design. As Another one is a fast inertial separator from FLOTU (Qi et al.
shown in Fig. 2 (f), this reactor consists of an annular riser and a [26]). This inertial separator was reported to separate more than
downer, connected smoothly at the top. The reason for choosing 96%–99% of the solids from the main stream within 0.05–0.3 s
an annular riser is to reduce the non-uniformity of flow structures (Wang et al. [27]). Considering that the residence time in a downer
in conventional risers and also to utilize the favorable heat and is only a fraction of a second, a cyclone with solids residence time
mass transfer behaviors in risers. When the reactions proceed to a in the order of 1–2 s is not acceptable.
certain extent, the reacting flow enters into the downer, where More recently, Wei et al. [28] devised a simpler gas–solids
over-cracking is prevented because of the plug-flow behavior in fast separator shown as Fig. 3 (c). This design makes the volume
downers. Feedstock can be injected at different positions along of the separator small, but the separation efficiency is very high
the riser or at the top of the downer, which depends on the specific (e.g., 98%).
applications. Residence times of gas and solids can be tuned in the
riser or in the downer. It is expected that this design provides a 4. Experimental study on multiphase flows in downer reactors
more flexible operation and control of fast reaction systems.
As discussed above, inlet design for downer reactors has The physics of gas–solids flows along the direction of gravity
drawn much attention because of its strong impact on the overall force imposes excellent hydrodynamics and mixing behavior
reactor performance. Although many designs have been studied on the overall performance of downer reactors. However, the
in the literature, more work has to be carried out for reliable potential applications of downer in fast competing reactions
designs in industry. would make the practical design very sensitive to slight changes
in the distinct developing flow structure in downer. As the
3.2. Outlet configuration for fast separation of gas and solids industrial concern and common interest in the special hydro-
dynamics phenomena, a large quantity of experimental studies
The gas–solids separator at downer outlet plays a similarly have been carried out in downer reactors.
significant role to the downer inlet design. The short, uniform
contact times between catalysts and gas require fast separation of 4.1. Developing flow phenomena in the downer column
gas and solids at the outlet to avoid consecutive reactions, e.g., the
excessive coke formation and over-cracking in FCC process. An The earlier researches revealed the axial flow structure of the
alternative could be to cool down the entire products flow very gas–solids flows in downer reactor, which have been well
quickly (i.e., quenching) to stop the reaction instantaneously. reviewed by Zhu et al. [5]. An overview of the one-dimensional

Fig. 4. Illustration of typical axial flow structure in downers.


Y. Cheng et al. / Powder Technology 183 (2008) 364–384 369

axial flow phenomena is sketched in Fig. 4. For detailed des- representative results measured in the big downer (418 mm i.d.
cription please refer to Zhu et al. [5]. and 7 m in height). These data came from the joint collaboration
More extensive studies on the developing fluid flow dyna- between FLOTU and PetroBras.
mics were carried out by Johnston et al. [19,20], Zhang et al.
[29,30], Zhang and Zhu [31], Lehner and Wirth [22]. The 4.2. Fully developed hydrodynamics and scale-up
developing hydrodynamics are generally related with the speci-
fic inlet designs. Cheng et al. [32] reported the experimental and At the beginning of the research on downers, the major
simulation results based on a simple inlet design with a center concern is to reveal the significantly different hydrodynamics in
solids jet. Wirth's group made great efforts to measure the solids the fully developed regions in riser and downer. It should be
fraction profiles along the flow development from the downer noted again that the only difference in operating a riser and a
inlet to the developed region near the downer exit by a novel X- downer is whether the flow direction is against (in riser) or
ray CT technique (Lehner and Wirth [22]; Grassler and Wirth along (in downer) the gravity force. However, the radial
[33]), as shown in Fig. 5. Typical distribution of a dense ring of profiles of hydrodynamic parameters, such as the solids fraction
solids near the wall was clearly disclosed, which agrees well and velocities, are greatly different between riser and downer.
with the experimental work by FLOTU. As summarized in Fig. 7, downer has rather uniform radial flow
Qian [34] conducted the measurements in a downer of 418- structure compared with riser. While, the unique feature of
mm i.d. with different gas nozzle arrangement. These experi- dense-ring distribution of solids in downer is still an academic
mental results are especially valuable to the industrial design interest and somewhat an unsolved problem in the physical
considering the large diameter of downer and the way to inject mechanism. Another aspect to point out is: on the contrary to
gas (e.g., feedstock in refinery process). Fig. 6 shows some the riser cases the solids velocity distribution in downer is very

Fig. 5. Solids fraction profiles along the flow development in downer (Grassler and Wirth, [33]).
370 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

However, the agglomerated particles in downer assemble


together very loosely and move downwards faster than the
dispersed particles and the gas in the fully developed region.
A more comprehensive study of radial flow structures in a
downer of 100 mm i.d. and 9.3 m in height was conducted in the
University of Western Ontario (Zhang et al. [29,30]; Zhang and
Zhu [31]). In their work, it was found that gas–solids flow in the
downer needs longer distance to reach fully developed. In the
beginning and middle parts of the downer, a dense annulus flow
structure was also found. At further downstream, the radial
profiles changed: the small peak near the wall disappeared. The
radial distribution of solids fraction showed dense wall – dilute
core structure at lower gas velocities but dense core – dilute
wall structure at higher gas velocities.
However, it is interesting to notice that the radial distribution
of solids fraction and solids velocity is dependent on the
diameter of the downer, i.e., the scale-up effect. Different shapes
of radial profiles of solids fraction and solids velocity were
found in smaller downer units. Herbert et al. [35] reported
parabolic profiles of solids velocities in the radial direction in a
50 mm i.d. downer. While, the solids fraction was really flat
along the radial direction.
Most of the experimental results obtained in downer diam-
eter greater than 100 mm show similar radial flow structures,
which were typically shown in the early work by FLOTU (see
for example, Wang et al. [27]), and Dr. Wirth's work using a X-
ray CT technique (Lehner and Wirth [36]). Most recently,
experimental study was done in a downer with inside diameter
of 418 mm. The scale-up effect was discussed in Zhang et al.
[37,38]. Fig. 8 summarizes the radial flow structures obtained
from different sized downers. The scale-up effect can be clearly
identified. However, how the wall properties including the
Fig. 6. Influence of gas nozzles at downer inlet on the distribution of solids physical properties of the wall and the diameter of the column
fractions in downer (Qian [34]). (a) 30-degree gas nozzle; (b) 45-degree gas nozzle. influence the radial flow structure in downers is still not clear in
physics.
consistent with the distribution of solids fraction. In riser, the To address the scale-up effect appropriately, scaling laws
clusters forming near the wall result in the slow motion of based on the dimensionless groups representing the physical
collected particles, and even cause the backflow along the wall. properties of phases, operating conditions, size of equipment

Fig. 7. Illustration of radial flow structure in the fully developed region of downer (experimental data from FLOTU in the column size of 140 mm i.d. for either riser and downer).
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 371

Fig. 8. Scale-up effect of fully developed hydrodynamics in downer reactors.

and so on may be essential to compare the different in these two reactors (Wei and Zhu [41]), which provided solid
hydrodynamics in downers with different dimensions. How- experimental evidence on the major advantage of downers over
ever, to establish such a methodology needs a large amount of riser, i.e., the approaching plug-flow behavior due to the effect
experimental data for validation. Under these circumstances, of flow direction. It can be further concluded that flow direction
proper modeling of multiphase flows would be more helpful to has the largest influence on the axial solids mixing in the
investigate the scale-up effect in details. vertical gas–solids suspension systems among many influen-
cing factors such as operating conditions, bed diameters,
4.3. Mixing and heat transfer particle properties and bed geometry.
Radial dispersion of solid particles was also measured using
Gas–solids mixing in downers was most studied by 1995, point sources of phosphorescent tracer technique and modeled
and has been well reviewed by Zhu et al. [5]. In spite of this, it is by a two-dimensional dispersion model (Wei et al. [39]). With
still interesting to mention the residence time distributions the increase of solids concentration, the dispersion of particle in
(RTDs) of solids measured by the novel phosphorescent tracer the radial direction is hindered. This result can be easily
technique. Wei et al. [39,40] obtained the solids RTDs in understood because the gas–solids flow in downers is more or
downer and riser, showing significantly different flow mechan- less like a segregated flow. Solid particles drop down with little
isms in these types of reactors. Narrow RTDs demonstrated the fluctuations in the radial movement. This flow phenomenon may
plug-flow behavior in downers. On the contrary, a bimodal become positive to improve the selectivity of a gas–solids
distribution of residence time with a long tail in risers was reaction system. However, the radial heat and mass transfer may
accurately measured, showing the fact of the strong backmixing be weak. This may be a major concern in designing a real downer
of solids in risers. Fig. 9 shows the typical solids RTDs reactor because the process design demands matching the rate of
measured in riser and downer and the calculated Peclet numbers heat transfer, mass transfer with fast chemical reactions.
372 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

Fig. 9. Typical solids RTDs measured in riser and downer and the calculated Peclet numbers in these two reactors (Wei and Zhu [41]).

Gas mixing results measured by using hydrogen/helium as transition (Cai et al. [47]; Chehbouni et al. [48]; Zijerveld et al.
the tracer gas showed that gas backmixing is very limited in the [49]), characterization of flow structures (Brereton and Grace
downer, but the lateral mixing is comparable with that in the [50]; Bai et al. [51]; Cheng et al. [52]), monitor of fluidization
riser (Wei et al. [42]; Cao and Weinstein [43]; Brust and Wirth quality (Schouten and van den Bleek [53]) and description of
[44]). cluster structures (Wei et al. [54]).
Heat transfer behavior is generally recognized to be However, transient dynamics in downers has not attracted
governed by the hydrodynamics in multiphase flows. Bed much attention compared with the similar research in other gas–
suspension density is the dominant factor influencing heat solids reactors. In risers, for example, clusters dominate the fluid
transfer since particle convection is mainly dependent on the dynamics, heat/mass transfer and mixing behavior. Therefore,
local bed density and solids velocity near the heat transfer study on this phenomenon would result in improved under-
surfaces. Investigations on heat transfer behaviors in downers standing on the flow mechanisms. Both direct measurement
were made (Ma and Zhu [21], [45]; Kim et al. [46]). Ma and
Zhu [45] measured the heat transfer coefficient between the
suspended surface and the gas-particle flow suspension in a
downer of 9.3 m high using a miniature cylindrical heat transfer
probe. The results demonstrated that bed suspension density is
the most influential factor. The average heat transfer coefficient
decreases with decreasing solids circulation rate due to the
decreased solids holdup. Whereas, the heat transfer coefficient
does not always decrease with increasing gas velocity given the
increased importance of gas convective heat transfer under high
gas velocities and low solids holdups in the downer. Heat
transfer coefficient is generally in the range of 100–150 W/
m2K, not significantly lower than what has been reported for
risers. The radial profiles of heat transfer coefficient are
consistent with the solids holdup profiles (see Fig. 10).

4.4. Micro-flow structure: clusters, statistical analysis, chaos


analysis

Two-phase flow structure is of great importance in under-


standing different gas–solids processes. This can be revealed in
a time-averaged manner or by looking at the transient dynamics.
Time-averaged study provided useful and straightforward
information to understand the steady flow fields. However, the
averaging procedure basically loses more information behind
the complex dynamics. Therefore, more advanced analysis on
the measured dynamics may lead to deeper understanding on the
underlying process. In fact, analyzing transient dynamics based
on the measured pressure and density fluctuations as well as
continuous images has been successfully applied in studying Fig. 10. Heat transfer coefficients in the downer (a) and in the riser (b) (Ma and
gas–solids reactors, which involves determination of regime Zhu [45]).
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 373

conditions. This implies that the transient moving behaviors of


particles in this region are similar from the point view of chaos
theory. On the other hand, there exist evidently different radial
distributions of dimension and entropy, which specifies the
difference of flow structures in radial direction for the riser and
downer. The local correlation dimension or Kolmogorov
entropy is in the same range of magnitude in both the riser
and downer and is strongly correlated with the local solids
fraction, which implies that the solids fraction is an important
factor that influences the turbulent structure in two-phase flow
systems.
Lu et al. [58] observed the clustering phenomena in a downer
of 0.09 m i.d. using a micro-video. The results showed that the
Fig. 11. A comparison of radial profiles of correlation dimension and
Kolmogorov entropy in the developed region of a riser and a downer (Cheng clusters in the downer were mainly in the form of floc and stick.
et al. [52]). The cluster size and frequency change along the bed height. The
video shots are shown in Fig. 12.

(Wei et al. [54]) and model simulation (Sunderasan [55]) 5. Modeling and simulation
illustrated the complex dynamics related with clusters. On the
contrary, large clusters may not exist in downers because of the The available experimental findings discovered the basic
downflow mechanism. Fast gas–solids flow in the same physical phenomenon in downers. However, this knowledge is
direction as gravity results in high shear force in the axial not enough to exploit these results to any new downer systems
direction so that the clusters are difficult to form transversely in because of the complexity in multiphase flows. The unknown
the radial direction. More likely, small, stream-like particle hydrodynamics is typically dependent upon scale, geometry,
aggregates may exist in downers, which comprises loosely physical properties of fluid(s) and solids, operating conditions,
packed particles. If so, reactor performance would not be etc. Theoretical modeling is therefore demanded for the
influenced by these clusters heavily. purposes of scale-up, reactor design and operation.
Tuzla et al. [56] studied the transient dynamics of local solids Based on the relatively uniform flow structures in downers,
concentrations in a downer of 15 cm i.d.. A needle capacitance Jin et al. [59] established a one-dimensional fluid flow model,
probe was used to obtain measurements of instantaneous solids which assumed the uniformity of flow structure in the radial
concentration in a small sampling volume of approximately direction. The gas–solids drag and gravity forces were taken
0.1 cm3. The measurements indicated that the local concentra- into account. The model successfully predicted the length of
tion did have excursions to significantly higher concentrations acceleration section of particles, and axial profiles of pressure &
for short durations. This behavior is indicative of clusters in the pressure drop, mean solids fraction and mean particle velocity.
downer. By an arbitrary threshold of 2 times standard deviation Kimm et al. [60] developed a hydrodynamic model for
above the mean value of the measured time series, the cluster laboratory scale downflow reactors, correlating published data
phase was defined. The results showed that for the roughly of solids holdup for fully developed flow by FLOTU
equivalent operating conditions, clusters in downers have researchers. The gas and particle velocity profiles in the radial
significantly shorter duration times than those in risers, by direction can be calculated. However, experimental data have
two orders of magnitude. It is seen that the actual time of cluster demonstrated various flow structures in downers with different
duration is relatively short, in the order of 0.1–10 ms. dimensions. This model would have very limited application.
Krol et al. [57] studied particle clustering in a down-flow Bolkan-Kenny et al. [61] compared riser and downer reactors
reactor of 5 cm i.d.. It is demonstrated that in down-flow for the fluid catalytic cracking (FCC) process based on a one-
systems that catalyst evolves, as strings of particles, with a dimensional hydrodynamic model incorporated with chemical
relatively large size distribution: average cluster sizes com- reactions in FCC process. A typical commercial reactor with
prised between 2 and 6 dp with most of these strings having 3.5 35 m height and 1 m diameter was adopted for both riser and
dp. It is shown that particle agglomeration is a very rapid downer. Although many advantages of downers over risers have
process taking place in downer units in a time scale much been claimed in the literature, the model predictions from this
smaller than the few seconds expected residence times. work only showed slightly more beneficial of downers over
Cheng et al. [52] introduced deterministic chaos theory to the risers. The authors attributed this to the absence of the proper
research of the transient dynamics, characterized by time series kinetic model for the FCC catalysts employed commercially
of local solids fraction, in downers. Special attention was and the low solids circulation rate used in the simulation.
focused on the comparison of dynamics in the developed region However, the omission of the radial non-uniformity and the
of a riser and a downer. Typical comparison is shown in Fig. 11. backmixing of gas and solids may be the cause since these two
The results showed correlation dimension (Dml) and Kolmo- factors have big influence on the reactor performance.
gorov entropy Kml) at the center are almost of the same values To consider the large difference in mixing behavior of gas
both in the riser and in the downer under the similar operating and solids in risers and downers, Wei et al. [62] developed a
374 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

Fig. 12. Images of solids clusters in the downer (Lu et al. [58]).

one-dimensional pseudo-homogeneous dispersion model. This less than 10 in riser and larger than 100 in downers. These
model combined a four-lump cracking kinetic model, hydro- reactor simulations suggested that because of the great
dynamic models for both riser and downer separately, and the reduction of axial gas backmixing in the downer, its yield of
dispersion mixing model, to predict the fluid catalytic cracking gasoline is much higher than that in the riser.
reaction performance. The results showed that axial gas With particular emphases, above models provided some
backmixing in the riser and the downer is a very important insights in understanding downer hydrodynamics and/or
factor that influences the yield of gasoline. When the axial downer reactors. However, the hydrodynamic models employed
Peclet number changes from 0.1 to 1000, the yield of gasoline are somewhat oversimplified so that they have little room to be
will increase approximately 11% under the same conversion, as extended for describing 2D or even 3D flow fields. Note that
shown in Fig. 13. Keep in mind that the axial Peclet numbers are hydrodynamics are strongly dependent on the geometry (e.g, at
inlet or outlet), and scale (e.g., downer diameter). To advance
the research and development of downers, a proper model based
on the physical mechanism is expected to give a theoretical
description and reasonable predictions of downer hydrody-
namics to contribute to the scale-up procedure and improvement
of the reactor design. Although downers are often claimed
having much more uniform flow structures than risers, the
unique radial structure, e.g., dense ring near the wall, found in
medium-size downers (e.g., 140 mm ID, 150 mm ID), imposes
difficulty in modeling and physical understanding, i.e., non-
monotonous profiles along the radial direction. This special
phenomenon is probably due to the mixed influence of gravity
force, wall effect and multiphase turbulence, which behave
differently in risers and in downers.
Computational fluid dynamics (CFD) is undergoing sig-
nificant expansion in terms of its applications in various
chemical processes. Basic mechanisms are included in the
governing equations, though a full understanding on multiphase
flows has not been achieved yet. To date, CFD calculations are
Fig. 13. Effect of axial dispersion on the gasoline yield in risers and downers increasingly being seen as an integral part of optimization or
(Wei et al. [62]). development plans for a much wider range of chemical
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 375

increased: the wall effect can be transferred to the center region


by the particle collisions.
Scale-up effect in the fully developed region of downers was
also investigated using the k–ɛ–Θ–kp model. As shown in Fig. 15,
in smaller size downers, e.g., less than 0.1 m i.d., the radial solids
distribution is fairly flat; the dense-ring flow structure only
appears in medium-size downers, such as 0.14 m i.d.; in much
larger downers (e.g., N 0.50 m ID), the solids fraction may show
monotonous trend along the radial direction with uniform core
region and very thin dense region at the wall. The simulation
results have very good consistency with the available experi-
mental data in the literature (e.g., Wang et al. [27]; Herbert et al.
[35]; Lehner and Wirth [36]; Zhang et al. [29,30], Zhang and Zhu
[31]; Zhang et al. [37,38]).
As stressed in Sections 2 and 3, design of gas and solids
distributor plays a dominant role in the reactor performance of
downers. Lots of efforts have been made in studying the hydro-
dynamics under various inlet structures. However, no criterion
could be followed in designing a new inlet geometry because of
the complicated multiphase contact in the entrance region, which
Fig. 14. CFD predictions of radial flow structures in the fully developed region is significantly structure-dependent. Under these circumstances,
of a downer (Cheng et al. [63]; Cheng [64]). experimental study is necessary but absolutely not economical to
compare different inlet designs. CFD simulation is a promising
processes. In particular, once they are accepted as a reasonably way to evaluate the initial designs of inlet geometry, at least for
accurate description of processes in a reactor they can be used a qualitative comparison at current stage. Cheng et al. [32]
for scale-up. simulated the gas–solids flows under a simple inlet geometry
Two types of CFD models, i.e., Lagrangian model and using the k–ɛ–Θ–kp model. As shown in Fig. 16, the predicted
Eulerian model, are being developed in parallel, in terms of the results obtained quantitative agreement with the experimental
different treatment on the discrete phase(s), in multiphase flow data for the solids jet process. Further analyses on the flow
simulations. Lagrangian models, or discrete particle models, development can be referred to Cheng et al. [32].
calculate the path of each individual particle with the second More recently, above CFD simulations based on a FORTRAN
law of Newton. While, Eulerian models treat the solid phase as a code were reproduced by those using commercial software, CFX
continuum and average out motion on the scale of individual 4.3 (Zheng et al. [65]). This is a very important step because CFX
particles, thus enabling computations by this method to treat provides a common platform to further develop a complete model
dense-phase flows of a realistic size. for downer reactors. With the help of CFX 4.3, simulation of heat
Under Eulerian–Eulerian scheme, Cheng et al. [63], [32]) transfer, mass transfer, reacting flow can be (easily) implemented,
developed a gas turbulence–solid turbulence model (k–ɛ–Θ–kp), especially under a complicated geometry.
taking into account the fast and dense gas–solids flow in downers.
The model comprises a k–ɛ turbulence model for gas phase, a kp
turbulence model and a kinetic theory description of solid stresses
characterized by granular temperature (Θ) for solid phase. Due to
the comprehensive consideration of gas–solid turbulent behavior
and the collisions between particles, the k–ɛ–Θ–kp model is
suitable for describing the gas–solids pipe flow in wide operating
ranges. The k–ɛ–Θ–kp model successfully predicted axial and
radial distributions of local solids fraction, local particle velocity
and pressure measured in a downer with 0.14 m i.d. and 7 m in
height. Quantitative agreement with experimental data was
obtained (e.g., Cheng et al. [63] and more details in Cheng
[64]). As shown in Fig. 14, at very dilute operations, i.e.,
Gs = 16 kg/m2s, the radial flow structures are very uniform,
without the dense-ring structure. With the increase of Gs or the
solids holdup in the system, the typical dense-ring flow appears
and sustains in wide operating ranges. The radial position
corresponding to the peak moves towards the center slightly with
the increase of Gs. This phenomenon may be attributed to the
strengthened wall effect when the solids concentration is Fig. 15. CFD predictions of scale-up effect of downers (Cheng et al. [63]).
376 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

Fig. 16. Model predictions of flow fields in the entrance region of a downer (Cheng et al. [32]).

However, it should be pointed out that modeling and 10 cm and the height of 10 m (to ensure the two-phase flow
simulation of downer hydrodynamics is not a solved problem at reaching fully developed) were simulated using the CFD–DEM
all. In terms of the basic physics, the radial profiles of solids method. The DEM code has been successfully incorporated
fraction and velocity in downer are very unique, which are NOT with the commercial CFD package of FLUENT 6.1. The
shown as a monotone relationship with the radial position. This predicted macro-scale flow structure had good agreement with
indicates that some extra physics may be incorporated to the the experiments. The distinct clustering phenomena at meso-
governing equations. By considering this, Cheng et al. [63,32] scale were revealed throughout the downer. Influences of the
expressed the influence of the wall on the hydrodynamics collision properties of the wall and the particles on the
through an empirical radial function of the coefficient of hydrodynamics in downer were investigated. Fig. 17 shows
restitution, which can predict the existence of peaks in the solids the flow development characterized by the particle distribution
fraction and velocity. If one directly adopts the default Eulerian at a transient. In the section of 0–3 m of the downer, the
models available in FLUENT or CFX, the simulation will not particles enter into the downer through five tubes, and then
give right predictions on downer hydrodynamics to our best
knowledge and experience. Jian and Ocone [66] noted that the
feature of these peaks resembles strong analogy with the
hydrodynamic behavior of Geldart A particles in riser. They
modified the solid stress tensor via the introduction of a counter-
diffusive term accounting for all inter-particle cohesive force,
together with the addition of gas-phase turbulence. The results
showed that the impact of gas-phase turbulence is not
significant. The counter-diffusive solid concentration term
plays the dominant role in the prediction capability.
Different from the Eulerian–Eulerian method, Eulerian–
Lagrangian model has clearer physics by considering the
particle–particle and particle–fluid contact at particle scale
using Newtonian laws. Discrete element method (DEM) is most
frequently applied in simulating the gas–solids flows in
fluidized beds. Zhao et al. [67] developed an up-to-date three-
equation linear dash-pot model for DEM to take full account of
the particle–particle interactions at the micro-scale, including
the normal contact force, the tangential contact force (including
static sliding friction force and sliding friction force) and the
moment (including the moment generated by the tangential
force, static rolling friction force and rolling friction force). The
gas–solids flows in a 2-dimensional downer with the width of Fig. 17. CFD–DEM simulation of hydrodynamics in downer (Zhao et al. [67]).
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 377

disperse laterally. Although the local solids fraction is relatively intact instead of becoming completely dispersed. This effect is
high in this section, clusters can be hardly observed. In the stronger at higher solids loading, since it is supposed that an
section of 3–4 m particles start to form clusters. And in the important factor influencing the dispersion process is the
section of 4–6 m the size and the number of clusters all momentum transfer from gas to solids. The solids stream moves
increases along the downer. In the rest section of 6–10 m, the through the reactor with a strongly reduced contacting
size of clusters still increases, but the number of clusters is efficiency compared to a situation with a homogeneous gas–
obviously reduced. From the instantaneous state of the bed by solids suspension. At lower solids loading the dispersion
simulation, the clusters show anomalous and are often in the process is fast and efficient. Consequently, agreement between
form of the floc structure. At the lower section of the bed, the model and experiment is good. Appropriate design of the
occurrence probability of clusters near the bed wall is higher downer inlet and application of sufficiently high gas velocities
than that in the bed center, and the size of clusters near the wall are expected to enable such a fast solids dispersion process at
is also larger, which agrees qualitatively with the observations high solids loading.
by a micro-video (Lu et al. [58]). Major contributions on hot downer experiments are from
Compared with the Eulerian–Eulerian model, the CFD– FLOTU and joint researches from the Petroleum Energy Center
DEM approach is more promising to explain the basic physics (PEC) in Tokyo and the King Fahd University of Petroleum and
in gas–solids flows in downers, e.g., the unique flow structure Minerals (KFUPM) in Dhahran. Common interest was focused
and scale-up effect, especially that the influence of walls and on the integration of refining and petrochemicals for greater
particle properties can be more reasonably modeled. FLOTU economic benefits from the conventional FCC process. It is
researchers are performing the extensive researches on this noted that FCC has contributed significantly in providing
topic, further involving the incorporation of heat transfer and propylene as the 2nd largest source of propylene, satisfying
reacting flow descriptions at particle scale. about one third of the world propylene demand. In China, a
workshop is organized annually by PetroChina to promote the
6. Process development to industrial demonstration discussion on developing new FCC processes to produce clearer
gasoline and light olefins as by-products.
Up to date, downer reactors have been investigated in several Abul-Hamayel [69] summarized their research work on the
fields such as high temperature catalytic reaction system, comparison of downer and riser based FCC process at high-
pyrolysis of coal and biomass, plasma reactor, etc. Prior to the severity conditions in pilot-plant experiments. Typical product
commercialization, more and more experiments at the lab-scale yields of conventional FCC and the two conceptual cases of HS-
and/or pilot-plant scale have been carried out. Encouraging FCC modes are listed in Table 1. Features of HS-FCC process
results in these hot model experiments demonstrated that thanks include high reaction temperature (above 550 °C), down-flow
to its plug-flow reactor performance downer reactor has unique reactor, short contact time, and high catalyst/oil ratio. Depend-
advantages when applied to the fast reaction systems with ing on the operating mode, the HS-FCC doubles the amount of
intermediates as desired products. light olefins and in another mode it provides three times more
light olefins accompanied with a minimum loss in gasoline. The
6.1. Fluid catalytic cracking (FCC) production of propylene is 2.1–3.6 times higher than the
conventional FCC process. HS-FCC gasoline has a high octane
Talman et al. [68] developed a downer reactor for the number and contains more heavy fractions. The content of
particular case of fluid catalytic cracking. Cracking of cumene olefins in the HS-FCC gasoline dropped by 50–85 wt.%
was chosen as reaction system for a first test series. At the very depending on the operating severity. By increasing the catalyst/
short contact times in the range of 60–400 ms, conversions up
to 70% have been determined. Among others, this is the result
of one of the main assets of the downer, viz. the fact that the
catalyst mass-flow rate is not limited like in riser reactors Table 1
(choking ratio). However, experimental data indicated that high Yields of conventional FCC and HS-FCC (Abul-Hamayel [69])
catalyst-to-oil ratios, up to 60, only result in high conversions Parameter Conventional HS-FCC
when the solids jet entering the downer reactor can be FCC
Case 1 Case 2
transformed into a homogeneously dispersed two-phase flow,
Reaction Temperatur (°C) 500 550 600
so that the catalyst is fully exposed to the gas mixture. A simple Conversion (wt.%) 75 87 90
plug-flow model in combination with the pseudo-homogeneous Product yield (wt.%)
approach was applied as a first attempt to simulate the behavior Ethylene 0.3 0.9 2.3
of the downer reactor. The model seems to describe the data Propylene 4.2 9.3 15.9
Butylene 5.6 12.2 17.4
reasonably well at low solids loading (or low catalyst/oil ratio)
Gasoline 53.6 49.5 37.8
in most cases, while the experimental values are substantially Light cycle oil (LCO) 17.6 8.8 6.6
lower than the calculated values at high solids loading Heavy cycle oil (HCO) 7.7 4.0 3.3
conditions. This can be explained qualitatively by the fact that Properties of gasoline (vol.%)
the solids dispersion at high solids loading is not homogeneous Olefins 13.5 9.6 5.1
Aromatics 28.0 37.0 37.0
and the compact solids stream entering the downer stays partly
378 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

oil ratio the effects of operating at high reaction temperature refer to their series work (Aitani et al. [16]; Maadhah et al. [70];
(thermal cracking) are minimized. Fig. 18 summarizes the Abul-Hamayel [71]; Abul-Hamayel et al. [72]; Abul-Hamayel
comparative result of downer and riser based FCC process at [69]; Abul-Hamayel et al. [73]).
high-severity operation in pilot-plant experiments (Abul- Based on more than 20 years' research on the downer
Hamayel [69]). It is observed that at same conversion, downer hydrodynamics, FLOTU established two sets of hot model
requires smaller catalyst-to-oil ratio, while obtains higher yield downer experimental apparatus shown in Fig. 19. Deng et al.
of gasoline, lower yields of coke and dry gas (see Fig. 18 (a–d)). [74] studied the deep catalytic cracking (DCC) process in a
Total yields of useful products (gasoline + light olefins) are downer reactor of 4.5 m in height and 13 mm in inner diameter
always higher in downer than in riser (Fig. 18 (e)). An (Fig. 19 (a)). The results showed that downer can significantly
interesting comparison is shown in Fig. 18 (f): at any constant improve the selectivity of desired products in comparison with
yield of total light olefins, downer offers more gasoline; at any the commercial riser adopting the same feed and catalyst: the
constant gasoline yield, downer offers more light olefins. These yields of propylene and gasoline increase by 3.71 and 7.30 wt.%,
results are very convincing to investigate larger-scale applica- respectively, while the yields of the dry gas and coke reduce by
tions of downers in FCC process. For detailed results please 6.77 and 1.98 wt.%. The comparison of the DCC process

Fig. 18. Comparison of downer and riser based FCC process at high-severity operation (Abul-Hamayel [69]).
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 379

Fig. 19. Hot downer experimental apparatus in FLOTU.

between the commercial riser and the hot downer is shown in from heavy feeds. The results showed that high olefins yields
Table 2. can be obtained from the DCP process: at the temperature of
In the same unit, Deng et al. [75] tested the downer catalytic 657 °C and residence time of 0.75 s, the total yields of ethylene,
pyrolysis (DCP) for a novel process for light olefins production propylene and butylene are up to 51.54 wt.%, while methane
and coke are suppressed. Typical comparison between experi-
Table 2
Comparison of DCC process between the commercial riser and the hot downer
mental results from DCC, DCP and HCC are given in Table 3.
(Deng et al. [74]) A newly built hot CFB apparatus including riser and downer
can be operated continuously shown in Fig. 19 (b). The downer
Operating conditions
is 14 mm diameter and 1500 mm in height; the riser is 14 mm
Units Commercial riser Downer experiment
diameter and 5000 mm in height. The feeding rate can be
Feeding temperature °C 380 350 adjusted from 200 to 1800 g/h. The inventory of catalyst is
Reaction temperature °C 565 606 about 4000–10,000 g. Reaction temperature can be operated at
Operating pressure Mpa 0.08 0.008
Catalyst/oil ratio kg/kg 10.87–11.80 37.54
300–720 °C and the regeneration temperature at 600–750 °C.
Residence time s 3.0–4.0 0.74 The pre-heating temperature is around 150–420 °C. RFCC
(residue FCC) gasoline upgrading was investigated in this unit.
Material balance The results are sketched in Table 4 (Li et al. [76]). The
Name Commercial riser Downer experiment upgrading results are very encouraging. It should be mentioned
Feed 100 100 that a series of experiments are still being carried out in FLOTU,
Products 1. Dry glass 11.03 4.26 especially to evaluate the catalysts for better use in downer
H2 S 0.20 / reactors, e.g., the catalysts with high activity and selectivity.
H2 0.29 0.13 More hot downer results will be reported in coming years.
CH4 3.39 0.91
As the first industrial demonstration of downer reactor in
C2H6 2.14 0.36
C2H4 5.01 2.86 refinery process, a circulating fluidized bed was built in Ji'nan
2. LPG 39.77 42.97 refinery (SINOPEC), shown in Fig. 20, in which the riser
C3H8 3.27 0.89 reactor is 500 mm in diameter and 40 m in height, while the
C3H6 19.54 23.25 downer is 600 mm in diameter and 15 m in height. The capacity
C4H10 2.84 3.46
is 150,000 t/a. Different from a single downer design, this
C4H8 14.12 15.37
3. Gasoline 21.47 28.77 demonstration unit adopted a patented technique (Jin et al.
4. LGO 19.23 17.82 [24]), i.e., coupled riser to downer (RtoD) reactor. The basic
5. Coke 8.5 6.52 idea is to utilize advantages of both riser and downer. The feed
6. Loss 0 3.67 oil first enters into riser, allowing for favorable initial phase
Summary 100 100
contact. For the high suspension density of catalysts in the riser,
380 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

Table 3
Comparison between experimental results from DCC-I, DCP and HCC (Deng
et al. [75])
Operating conditions
Unit DCC-1 DPC HCC
Feed VGO Deasphalted oil Atmospheric
residue
Catalyst CPR-1 CHP-1 LCM-5
Reactor Riser + bed Downer Riser
Feeding temperature 350
Cracking temperature °C 550 659 700
Operating pressure Mpa 0.005
Catalyst/oil ratio kg/kg 112 38.64 19.6
Residence time s 0.75 1.94

Material balance
Name DCC-1 DPC HCC
Feed 100 100 100
Products 1. Gas 46.03 61.22 59.57
H2 0.45
CH4 9.16
C2H6 2.74
C2H4 3.76 16.06 22.94 Fig. 20. Photo of the industrial demonstration of downer reactor (Ji'nan
C3H8 0.80 Refinery, SINOPEC).
C3H6 17.41 24.57 13.30
C4H10 0.99
C4H8 13.39 10.91 5.85
Two series of experiments were tested in this demonstration unit.
2. Gasoline 25.12 14.27 29.70
3. LCO 20.83 12.74 • The first one tested the feed of VGO plus DAO for deep
4. Coke 7.61 9.54 9.97 catalytic cracking. Compared with the DCC process in the
5. Loss 0.41 − 2.23 0.76 riser, downer DCC process suppressed the dry gas forma-
Summary 100 100 100
tion by 2.27 wt.%. While, the yield of gasoline increased
1.06 wt.%, diesel oil increased 8.38 wt.% and propylene
conversion can be ensured. Then the reaction mixture enters 0.75 wt.%. The overall yield of liquid products increased
downer, where favorable RTDs, plug-flow pattern and uniform 0.27 wt.%. It should be mentioned that the catalyst was not
flow structure exist. Accordingly, over-cracking is limited and active enough in the first try. It is expected that the industrial
selectivity is improved. This coupled reactor design also makes tests should have better results if the active catalysts are used.
the operation very flexible: the feed positions can be adjusted • The second series of experiment was done using residue oil for
either along the riser or along the downer and feeds can be RFCC process. The yields of LPG and propylene increased by
different at different feeding points. 11.45 and 5.06 wt.%, respectively. While, the gasoline and
diesel oil yields decreased by 4.45 and 7.03 wt.%, respectively.
Table 4 The dry gas decreased 1.32 wt.%. In the two series of
RFCC gasoline upgrading in a hot downer reactor (Li et al. [76]) experiments, coke has little change in yield. Promising results
Operating conditions directed to the cleaner gasoline production with the olefin
Units Exp. 1 Exp. 2 content in gasoline of 30.7%, much less than the result from
Reaction temperature °C 611 693
conventional riser RFCC using the same feed and catalyst. In
Catalyst/oil ratio – 21.9 24.9 addition, the octane number (RON) of downer RFCC gasoline
Gas residence time s 0.51 0.41 maintained high, i.e., 94.8.

Material balance Above results would greatly encourage the further industrial
Weight fraction (%) Feed oil Exp. 1 Exp. 2 application of downer reactors. The full commercialization of
Paraffins 34.57 31.19 21.67 downer FCC refinery is expected to come soon.
Olefins 44.00 8.41 7.95
Naphthenes 7.98 10.01 4.96 6.2. Pyrolysis of biomass and coal
Aromatics 8.10 44.80 62.04
Others – 5.56 3.32
Octane number (MON/RON) 78.5/88.8 80.2/91.4 83.5/96.8 Berg et al. [77] proposed an ultra-rapid fluidized reactor,
Yields (%) which has now been successfully applied to biomass pyrolysis.
LPG 33.43 35.99 It typically consists of a “mixer”, downer and a gas–solid
Gasoline 59.02 41.51 separator. The main products are liquid substances, which can
C2H4 + C3H6 + C4H8 24.46 33.38
be utilized as fuel or chemical product.
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 381

To widen the application of the downer reactor, Kim et al. the downer reactor unique advantages for the potential
[78] developed a new type of coal gasifier, namely the downer applications in the ultra-short contact time processes, especially
gasifier. The heat is supplied from coal combustion in a riser that the high-severity operation can be implemented in downer.
reactor. The circulation of solids within the reactor loop provides Key features of downer reactors are summarized as below:
energy transfer from the combustion zone (riser) to the
gasification zone (downer, gas separator and loop-seal) with (1) Due to the co-current flow in the direction of gravity,
higher carbon conversions. In the present gasifier, the coal fed there is no minimum gas velocity to suspend particles. As
from the top was converted by the following reactions: a result, the solids/gas load ratio can be high.
combustion and gasification reactions of circulating char in the (2) Both axial and radial gas–solids flow structures are rela-
downer, coal pyrolysis, secondary and homogeneous gas-phase tively uniform, resulting in well-proportioned gas–solids
reaction in the gas–solid separator, volatile and char combustion contact time. Localized high temperature regions are
in the loop-seal, secondary and homogeneous gas-phase reaction unlikely to exist due to the efficient gas–solids contact.
of the product gas in the freeboard region, and combustion (3) Axial backmixing of solids is effectively reduced. How-
reaction of circulating reaction in the riser. Experiments were ever, the radial mixing of solids may need to be enhanced
carried out in a downer of 0.1 m i.d. and 5.0 m high. The reaction together with the heat transfer.
temperature is in the range of 750–850 °C. The results showed (4) The inlet design has great impact on the downer perfor-
that by changing the reactant gas supplied into the loop-seal for mance, considering the fact that the overall residence time
solids circulation from steam to air, product gas yield and carbon is short (e.g., from milliseconds to seconds) so that the
conversion increase, whereas calorific value of the product gas well distribution of solids at injection in such a short time
decreases with reaction temperature. is challenging.
A downer reactor with the highest severity operation would be (5) The outlet design plays an important role in the reactor
the coal pyrolysis in thermal plasma for acetylene. This process performance by the demand of fast separation and/or
opens up a direct means for producing acetylene, together with the quenching of the multiphase reacting flow.
valuable carbon materials in solid state and some ethylene, (6) As a novel FCC reactor, downer typically obtains more
methane, hydrogen and carbon monoxide as the secondary yields of middle distillates (e.g., gasoline and LPG) and less
products. The process essentially consists of several steps: (1) the yields of coke and dry gas than the conventional riser. Under
design of a hydrogen plasma torch, (2) fast heating of coal high-severity operation, the yields of light olefins especially
powders in milliseconds to release the volatile matter in coal in a propylene are increased, which results in an efficient inte-
downer reactor, (3) fast quenching of the reacting stream to gration of refining and petrochemical processes.
prevent the decomposition of acetylene into soot and hydrogen.
Since the complex reaction process is operated at very high- Based on the great efforts made in the R&D activities on
severity conditions, e.g., reaction temperature at 1500 K to downer reactors, it is approaching to the time when to
~3000 K, high gas velocity and residence time in milliseconds, commercialize downer as a new-generation FCC reactor.
the reactor design meets great challenges such as the nozzle design Successful industrial demonstration with the capacity of
for coal injection into the hot hydrogen stream, severe coking 150 kt/a of heavy oil refining in China has paved the way to
problems which prohibit the continuous operation, etc. So far, this larger-scale production at ~ 1 Mt/a. As for the coal pyrolysis in
process has not been commercialized since the early proposals by hydrogen plasma for acetylene, a 5 MW plant is being built in
Bond et al. [79], Nicholson and Littlewood [80]. Several China, which will direct to a new route to the petrochemical
companies had been working on pilot-plant experiments on coal processes from a clean coal utilization.
pyrolysis (e.g., Patrick and Gannon [81]; Bitter et al. [82,83]), but With deepened understanding on downer reactors, new idea on
it is hard to find the recent update on this process development. reactor coupling technique has been proposed to take full advan-
China started to investigate the coal pyrolysis in late 1990s. To tage of downer and another reactor (e.g., riser, turbulent bed, etc.) to
date, encouraging progress has been made by the joint efforts from implement multi-functional purposes such as to adjusting products
Xinjiang Tianye Group, Chinese Academy of Sciences, Fudan distribution in multi-zones for a complex reaction processes (Jin et
University and Tsinghua University (FLOTU). A 2 MW thermal al. [24]; Wei et al. [84]). This will help open new chances to apply
plasma torch was successfully built and run well with hydrogen as downer-coupled reactors into some specific processes, e.g., flexible
the heat carrier and reacting atmosphere. The cost for producing FCC, chemical looping technique (e.g., combustion or reforming),
1 kg acetylene is about 10 kWh, which should be even reduced and multi-zone polymerization reactor technique.
after careful optimization of process conditions and reactor In addition to the new process development of downer
design, e.g., nozzle design for coal injection and quenching mode. reactors, fundamental studies are still the core in academia to
understand the unique multiphase flow phenomena. In
7. Conclusions particular, following aspects may receive more attractions,
involving high-density downer and related issues (e.g., Liu et al.
A comprehensive review on the fundamental studies and [85]; Li et al. [86]; Chen and Li [87]), scale-up methodology,
industrial development at pilot-scale of downer reactors was multiphase flow modeling and simulation incorporating with
made. It has been acknowledged that downer has favorable flow reacting flows, measurement techniques to capture fast flow
structures and plug-flow reactor performance, which endowed phenomenon and fast reaction process in kinetics.
382 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

List of symbols gives special thanks to Dr. J.R Grace, Dr. L.-S. Fan, Dr. J.X.
dp mean particle diameter, m Zhu, Dr. X.T. Bi, Dr. P. Cai and Dr. D.R. Bai for their great
D diameter of downer/riser, m support in nominating Prof. Jin as the candidate for 2006
Dml local correlation dimension, (−) Lectureship Award in Fluidization (PSRI).
FD drag force between gas and particles, N
FG gravitational force, N References
Gs solids circulation rate, kg/m2s
h heat transfer coefficient, W/m2K [1] J.R. Grace, Contacting modes and behaviour classification of gas–solid
and other two-phase suspensions, Can. J. Chem. Eng. 64 (1986) 353–363.
H total height of dower/riser, m
[2] J.R. Grace, High-velocity fluidized bed reactors, Chem. Eng. Sci. 45 (1990)
Kml Kolmogorov entropy, bits/s 1953–1966.
P pressure in downer/riser, Pa [3] H.T. Bi, J.R. Grace, Flow regime diagrams for gas–solid fluidization and
Pe Peclet number, (−) upward transport, Int. J. Multiphase Flow 21 (1995) 1229–1236.
r radial position, m [4] K.S. Lim, J.X. Zhu, J.R. Grace, Hydrodynamics of gas–solid fluidization,
Int. J. Multiphase Flow 21 (1995) 141–193.
R radius of downer/riser, m
[5] J.X. Zhu, Y. Jin, Z.Q. Yu, J.R. Grace, A. Issangya, Cocurrent downflow
Ug superficial gas velocity, m/s circulating fluidized bed (downer) reactors — a state of the art review,
Vg gas velocity, m/s Can. J. Chem. Eng. 73 (1995) 662–677.
V̄ g cross-sectionally averaged gas velocity, m/s [6] M. Kwauk, J. Li, D. Liu, Particulate and aggregative fluidization —
Vp particle velocity, m/s 50 years in retrospect, Powder Technol. 111 (2000) 3–18.
[7] L. Reh, Challenges of circulating fluid-bed reactors in energy and raw
V̄ p cross-sectionally averaged particle velocity, m/s
materials industries, Chem. Eng. Sci. 54 (1999) 5359–5368.
x distance from the inlet of downer, m [8] R.J. Gartside, QC — a new reaction system, in: J.R. Grace, L.W. Shemilt,
z distance from the inlet of downer, m M.A. Bergougnou (Eds.), Fluidization VI, 1989, pp. 25–32, Engineering
Foundation, New York.
Greek symbols [9] B. Gross, M.P. Ramage, FCC reactor with a downflow reactor riser, U.S.
Patent 4385985, 1983.
ε voidage, (−)
[10] J.R. Murphy, Evolutionary design changes mark FCC process, Oil Gas J.
ε̄ g cross-sectionally averaged voidage, (−) 18 (1992) 49–58.
ε̄ s cross-sectionally averaged solids holdup, (−) [11] B. Gross, Heat balance in FCC process and apparatus with downflow
Θ granular temperature, m2/s2 reactor riser, U.S. Patent 4411773, 1983.
[12] P.K. Niccum, D.P. Bunn, Catalytic cracking system, U.S. Patent 4514285,
1985.
List of acronyms
[13] T.S. Dewitz, Downflow fluidized catalytic cracking system, U.S. Patent
CFD computational fluid dynamics 4797262, 1989.
CSTR continuous stirred tank reactor [14] G.P. Muldowney, FCC process with upflow and downflow reactor, U.S.
DAO de-asphalted oil Patent 5468369, 1995.
DCC deep catalytic cracking process [15] R. Pontier, F. Hoffmann, P. Galtier, Downflow fluid catalytic cracking
process and apparatus, U.S. Patent 5344554, 1994.
DCP downer catalytic pyrolysis process
[16] A. Aitani, T. Yoshikawa, T. Ino, Maximization of FCC light olefins by high
DEM discrete element method severity operation and ZSM-5 addition, Catalysis Today 60 (2000) 111–117.
FCC fluid catalytic cracking [17] C.M. Qi, Z.Q. Yu, Y. Jin, D.R. Bai, W.H. Yao, Hydrodynamics of cocurrent
FLOTU Fluidization Lab of Tsinghua University downwards fast fluidization (I) and (II), J. Chem. Ind. Eng. 41 (1990)
HCC heavy oil contact cracking process 273–290 (in Chinese).
[18] C.S. Cao, Y. Jin, Z.Q. Yu, Z.W. Wang, The gas–solids velocity profiles and
HSCC high-severity catalytic cracking process
slip phenomenon in a concurrent downflow circulating fluidized bed, in:
KFUPM King Fahd University of Petroleum and Minerals A.A. Avidan (Ed.), Circulating Fluidized Bed Technology IV, AIChE, New
MON Motor Octane Number York, 1994, pp. 406–413.
PEC Petroleum Energy Center [19] P.M. Johnston, H.I. de Lasa, J.X. Zhu, Axial flow structure in the entrance
PSRI Particulate Solids Research Inc. region of a downer fluidized bed: effects of the distributor design, Chem.
Eng. Sci. 54 (1999) 2161–2173.
PTF Particle Technology Forum
[20] P.M. Johnston, J.X. Zhu, H.I. de Lasa, H. Zhang, Effect of distributor
RFCC residual oil fluid catalytic cracking process designs on the flow development in downer reactor, AIChE J. 45 (1999)
RON Research Octane Number 1587–1592.
RTDs residence time distributions [21] Y. Ma, J.X. Zhu, Characterizing gas and solids distributors with heat
RtoD coupled riser to downer reactor transfer study in a gas–solids downer reactor, Chem. Eng. J. 72 (1999)
235–244.
VGO vacuum gas oil
[22] P. Lehner, K.E. Wirth, Effects of the gas solids distributor on the local and
overall solids distribution in a downer reactor, Can. J. Chem. Eng. 77
Acknowledgements (1999) 199–206.
[23] C.L. Briens, C. Mirgain, M.A. Bergougnou, M. DelPozo, R. Loutaty,
The authors would like to thank the financial supports Evaluation of gas–solids mixing chamber through cross correlation and
Hurst's analysis, AIChE J. 43 (1997) 1469–1479.
from NSFC, SINOPEC, PetroChina and PetroBras on the
[24] Y. Jin, F. Wei, Y. Cheng, Z.W. Wang, J.F. Wang, Gas and Solid Parallel-
downer-related projects. Recent financial supports from flow Folding Type Quick Fluidized-bed Reactor, 2000, CN 1113689C.
NSFC (No. 20306012), FANED (No. 200245), and SRFDP [25] R.J. Gartside, H.N. Woebecke, Solids feeding device and system, U.S.
(No. 20050003028) are greatly acknowledged. Prof. Yong Jin Patent 4338187, 1982.
Y. Cheng et al. / Powder Technology 183 (2008) 364–384 383

[26] C.M. Qi, Z.Q. Yu, Y. Jin, X.L. Cui, X.X. Zhong, A novel inertial separator [53] J.C. Schouten, C.M. van den Bleek, Monitoring the quality of fluidization
for gas–solid suspension in concurrent downflow circulating fluidized using the short-term predictability of pressure fluctuations, AIChE J. 44
beds, Pet. Refin. 20 (1989) 51–56 (in Chinese). (1998) 48–60.
[27] Z.W. Wang, D. Bai, Y. Jin, Hydrodynamics of cocurrent downflow [54] F. Wei, G.Q. Yang, Y. Jin, Z.Q. Yu, Characteristics of cluster in a high
circulating fluidized bed, Powder Technol. 70 (1991) 271–275. density circulating fluidized bed, Can. J. Chem. Eng. 73 (1995) 650–655.
[28] F. Wei, Y. Jin, Z. Qian, Y.H. Yang, Z.W. Wang, Wall-attached Cutting Type [55] S. Sundaresan, Modeling the hydrodynamics of multiphase flow reactors:
Fast Gas–solid Separator, 2000, CN 1267564. current status and challenges, AIChE J. 46 (2000) 1102–1105.
[29] H. Zhang, J.X.J. Zhu, M.A. Bergougnou, Flow development in a gas– [56] K. Tuzla, A.K. Sharma, J.C. Chen, T. Schiewe, K.E. Wirth, O. Molerus,
solids downer fluidized bed, Can. J. Chem. Eng. 77 (1999) 194–198. Transient dynamics of solid concentration in downer fluidized bed, Powder
[30] H. Zhang, J.X. Zhu, M.A. Bergougnou, Hydrodynamics in downflow Technol. 100 (1998) 166–172.
fluidized beds (1): solids concentration profiles and pressure gradient [57] S. Krol, A. Pekediz, H. de Lasa, Particle clustering in down flow reactors,
distributions, Chem. Eng. Sci. 54 (1999) 5461–5470. Powder Technol. 108 (2000) 6–20.
[31] H. Zhang, J.X. Zhu, Hydrodynamics in downflow fluidized beds (2): particle [58] X.S. Lu, S.G. Li, L. Du, J.Z. Yao, W.G. Lin, H.Z. Li, Flow structures in the
velocity and solids flux profiles, Chem. Eng. Sci. 55 (2000) 4367–4377. downer circulating fluidized bed, Chem. Eng. J. 112 (2005) 23–31.
[32] Y. Cheng, F. Wei, Y.C. Guo, Y. Jin, CFD simulation of hydrodynamics in [59] Y. Jin, Z.Q. Yu, D.R. Bai, C.M. Qi, X.X. Zhong, Modeling of gas and
the entrance region of a downer, Chem. Eng. Sci. 56 (2001) 1687–1696. solids two-phase flow for cocurrent downflow gas–solid suspensions,
[33] T. Grassler, K.E. Wirth, X-Ray computer tomography — potential and Chem. Reaction Eng. Technol. 6 (1990) 17–23 (in Chinese).
limitation for the measurement of local solids distribution in circulating [60] N.K. Kimm, F. Berruti, T.S. Pugsley, Modeling the hydrodynamics of
fluidized beds, in: T. York, T. Dyakowski, T. Peyton, A. Hunt (Eds.), downflow gas–solids reactors, Chem. Eng. Sci. 51 (1996) 2661–2666.
Proceedings of the 1st World Congress on Industrial Process Tomography: [61] Y.G. Bolkan-Kenny, T.S. Pugsley, F. Berruti, Computer simulation of the
Buxton (Greater Manchester), 1999, pp. 402–409. performance of fluid catalytic cracking (FCC) risers and downers, Ind.
[34] Z. Qian, Experimental investigation on the hydrodynamics and inlet outlet Eng. Chem. Res. 33 (1994) 3043–3052.
structure in a large scale downer, Ph.D. Thesis, Tsinghua University, 2006. [62] F. Wei, X. Ran, R.J. Zhou, G.H. Luo, Y. Jin, A dispersion model for fluid
[35] P.M. Herbert, T.A. Gauthier, C.L. Briens, M.A. Bergougnou, Application catalytic cracking riser and downer reactors, Ind. Eng. Chem. Res. 36
of fiber optic reflection probes to the measurement of local particle velocity (1997) 5049–5053.
and concentration in gas–solid flow, Powder Technol. 80 (1994) 243–252. [63] Y. Cheng, Y.C. Guo, F. Wei, Y. Jin, W.Y. Lin, Modeling the hydrodynamics of
[36] P. Lehner, K.E. Wirth, Characterization of the flow pattern in a downer downer reactors based on kinetic theory, Chem. Eng. Sci. 54 (1999) 2019–2027.
reactor, Chem. Eng. Sci. 54 (1999) 5471–5483. [64] Y. Cheng, Numerical simulation and nonlinear dynamic analysis of gas–
[37] M.H. Zhang, Z. Qian, H. Yu, F. Wei, The solid flow structure in a cir- solid two-phase flow, Ph.D. Thesis, Tsinghua University, 2000.
culating fluidized bed riser/downer of 0.42-m diameter, Powder Technol. [65] Y. Zheng, Y. Cheng, F. Wei, Y. Jin, CFD simulation of hydrodynamics in
129 (2003) 46–52. downer reactors, Chem. Eng. Comm. 189 (2002) 1598–1610.
[38] M.H. Zhang, Z. Qian, H. Yu, F. Wei, The near wall dense ring in a large- [66] H. Jian, R. Ocone, Modeling the hydrodynamics of gas–solid suspension
scale down-flow circulating fluidized bed, Chem. Eng. J. 92 (2003) in downers, Powder Technol. 138 (2003) 73–81.
161–167. [67] Y.Z. Zhao, Y. Cheng, Y.L. Ding, Y. Jin, Understanding the hydrodynamics
[39] F. Wei, Z.W. Wang, Y. Jin, Z.Q. Yu, W. Chen, Dispersion of lateral and in a 2D downer by CFD–DEM simulation, in: F. Berruti, X.T. Bi, D.
axial solids in a cocurrent downflow circulating fluidized bed, Powder Puglsey (Eds.), Fluidization XII, 2007, pp. 855–862, Engineering
Technol. 81 (1994) 25–31. Foundation, Vancouver, Canada.
[40] F. Wei, Y. Jin, Z.Q. Yu, W. Chen, S. Mori, Lateral and axial mixing of the [68] J.A. Talman, R. Geier, L. Reh, Development of a downer reactor for fluid
dispersed particles in CFB, J. Chem. Eng. Jpn. 28 (1995) 506–510. catalytic cracking, Chem. Eng. Sci. 54 (1999) 2123–2130.
[41] F. Wei, J.X. Zhu, Effect of flow direction on axial solid dispersion in gas [69] M.A. Abul-Hamayel, Comparison of downer and riser based fluid catalytic
solids cocurrent upflow and downflow systems, Chem. Eng. J. 64 (1996) cracking process at high severity condition: a pilot plant study, Petrol. Sci.
345–352. Technol. 22 (2004) 475–490.
[42] F. Wei, Y. Jin, Z.Q. Yu, J.Z. Liu, Gas Mixing in cocurrent downflow [70] A.G. Maadhah, M. Abul-Hamayel, A.M. Aitani, T. Ino, T. Okuhara,
circulating fluidized bed, Chem. Eng. Technol. 18 (1995) 59–62. Down-flowing FCC reactor — increases propylene, gasoline make Oil Gas
[43] C.S. Cao, H. Weinstein, Characterization of downflowing high velocity J. 98 (2000) 66–70.
fluidized beds, AIChE J. 46 (2000) 515–522. [71] M.A. Abul-Hamayel, Effect of feedstocks on high-severity fluid catalytic
[44] H. Brust, K.E. Wirth, Residence time behavior of gas in a downer reactor, cracking, Chem. Eng. Technol. 25 (2002) 65–70.
Ind. Eng. Chem. Res. 43 (2004) 5796–5801. [72] M.A. Abul-Hamayel, M. Abdul-Bari Siddiqui, T. Ino, A.M. Aitani,
[45] Y.L. Ma, J. Zhu, Heat transfer in the downer and the riser of a circulating Experimental determination of high-severity fluidized catalytic cracking
fluidized bed — a comparative study, Chem. Eng. Technol. 24 (2001) (HS-FCC) deactivation constant, Appl. Catal., A Gen. 237 (2002) 71–78.
85–90. [73] M.A. Abul-Hamayel, A.M. Aitani, M.R. Saeed, Enhancement of
[46] Y.J. Kim, J.H. Bang, S.D. Kim, Bed-to-wall heat transfer in a downer propylene production in a downer FCC operation using a ZSM-5 additive,
reactor, Can. J. Chem. Eng. 77 (1999) 207–212. Chem. Eng. Technol. 28 (2005) 923–929.
[47] P. Cai, Y. Jin, Z.Q. Yu, Z.W. Wang, Mechanism of flow regime transition [74] R.S. Deng, F. Wei, Y. Jin, Q.H. Zhang, Y. Jin, Experimental study of the
from bubbling to turbulent fluidization, AIChE J. 36 (1990) 955–956. deep catalytic cracking process in a downer reactor, Ind. Eng. Chem. Res.
[48] A. Chehbouni, J. Chaouki, C. Guy, D. Klvana, Characterization of the flow 41 (2002) 6015–6019.
transition between bubbling and turbulent fluidization, Ind. Eng. Chem. [75] R.S. Deng, F. Wei, Y. Jin, Q.H. Zhang, Y. Jin, Downer catalytic pyrolysis
Res. 33 (1994) 1889–1896. (DCP): a novel process for light olefins production, Chem. Eng. Technol.
[49] R.C. Zijerveld, F. Johnsson, A. Marzocchella, J.C. Schouten, C.M. van den 25 (2002) 711–716.
Bleek, fluidization regimes and transitions from fixed bed to dilute [76] Q. Li, F. Wei, G.H. Luo, L. Wang, Q.H. Zhang, Catalytic conversion of
transport flow, Powder Technol. 95 (1998) 185–204. FCC gasoline in a downer reactor, Petrochem. Technol. 33 (2004)
[50] C.M.H. Brereton, J.R. Grace, Microstructural aspects of the behavior of 402–406 (in Chinese).
circulating fluidized bed, Chem. Eng. Sci. 48 (1993) 2565–2572. [77] D.A. Berg, C.L. Briens, M.A. Bergougnou, Reactor development for the
[51] D. Bai, E. Shibuya, Y. Masuda, N. Nakagawa, K. Kato, Flow structure in a ultrapyrolysis reactor, Can. J. Chem. Eng. 67 (1989) 96–101.
fast fluidized bed, Chem. Eng. Sci. 51 (1996) 957–966. [78] Y.J. Kim, S.H. Lee, S.D. Kim, Coal gasification characteristics in a downer
[52] Y. Cheng, F. Wei, Q. Lin, Y. Jin, A comparison of local chaotic behaviors reactor, Fuel 80 (2001) 1915–1922.
in a riser and a downer, in: L.-S. Fan, T.M. Knowlton (Eds.), Fluidization [79] R.L. Bond, I.F. Galbraith, W.R. Ladner, G.I.T. McConnell, Production of
IX, 1998, pp. 613–620, Engineering Foundation, New York. acetylene from coal, using a plasma jet, Nature 200 (1963) 1313–1314.
384 Y. Cheng et al. / Powder Technology 183 (2008) 364–384

[80] R. Nicholson, K. Littlewood, Plasma pyrolysis of coal, Nature 236 (1972) [84] F. Wei, G.H. Luo, Z.Q. Li, Y. Jin, A Novel Coupled Downer–riser Reactor
397–400. and Technology for Catalytic Cracking, 2002, CN 1162514C.
[81] A.J.J. Patrick, R.E. Gannon, A 1 MW prototype arc reactor for processing [85] W. Liu, K.B. Luo, J.X. Zhu, J.M. Beeckmans, Characterization of high-
coal to chemicals, in: P.N. Cheremisinoff, O.G. Farah, R.P. Ouellette density gas–solids downward fluidized flow, Powder Technol. 115 (2001)
(Eds.), Radio Frequency/Radiation and Plasma Processing: Industrial 27–35.
Applications & Advances, Technomic Pub Co., Lancaster, 1985. [86] Z.Q. Li, C.N. Wu, F. Wei, Y. Jin, Experimental study of gas–solids flow in
[82] D. Bittner, H. Baumann, Relation between coal properties and acetylene a new coupled circulating fluidized bed, Powder Technol. 139 (2004)
yield in plasma pyrolysis, Fuel 64 (1985) 1370–1374. 214–220.
[83] D. Bittner, W. Wanzl, The significance of coal properties for acetylene [87] H.Z. Chen, H.Z. Li, Characterization of a high-density downer reactor,
formation in a hydrogen plasma, Fuel Process. Technol. 24 (1990) 311–316. Powder Technol. 146 (2004) 84–92.

View publication stats

You might also like