You are on page 1of 9

Chemical Engineering Journal 334 (2018) 1383–1391

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Enhancing anaerobic fermentation performance through eccentrically T


stirred mixing: Experimental and modeling methodology
Yuankai Huanga, Farzaneh MahmoodPoor Dehkordya, Yan Lia, Sharareh Emadib,

Amvrossios Bagtzogloua, Baikun Lia,
a
Department of Civil and Environmental Engineering, University of Connecticut, Storrs, CT 06269, United States
b
Department of Biomedical Engineering, University of Connecticut, Storrs, CT 06269, United States

G RA P H I C A L AB S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: Anaerobic hydrogen production (AHP) is a biochemical process capable of producing hydrogen and fatty acids
Anaerobic hydrogen production (AHP) from organic substrates in acidogenic phase. Currently, almost all of anaerobic systems use centrally stirred
Eccentrically stirred mixing mixing where the flow is essentially rotated uniformly in the radial direction. Given the high energy requirement
Mixing eccentricity of mixing and the urgent needs for enhancing the contact between microbial particles and substrates, a new type
Mixing height
of mixing strategy should be developed. This study focused on the enhancement of biogas production and or-
Lyapunov exponent
ganic removal efficiency in AHP systems through a novel chaotic mixing regime. The stirring impeller was
Navier-Stokes equations
operated eccentrically with the stirring center eccentrically located 2 and 4 cm off the center, respectively and at
different heights (5, 10.5 and 16 cm). The eccentrically stirred mixing was expected to enhance the mixing
efficiency and accelerate the contact between biomass and organic substrates in the AHP system. A numerical
finite element model was developed to simulate the flow patterns and the biomass distribution, through which
the Lyapunov exponent λ values were calculated in order to assess the strength of eccentrically stirred mixing at
different mixing eccentricities. The model simulations showed that at the height of 10.5 and 16 cm, the
Lyapunov exponent λ increased as the mixing eccentricity increased. This corresponded well with the experi-
mental results, which showed that the organic removal efficiency and biogas production increased with the
mixing eccentricity. Compared with centrally stirred mixing, eccentrically stirred mixing enhanced the inter-
action of flow and biomass particles and ultimately enhanced the biogas production and organic removal in AHP
systems.

1. Introduction capable of producing hydrogen and organic acids (e.g., acetic acid,
butyric acid) from organic substrates (e.g., glucose and wastewater)
Anaerobic hydrogen production (AHP) is a biochemical process under acidogenic phase. Compared with methane production from


Corresponding author.
E-mail address: baikun.li@uconn.edu (B. Li).

https://doi.org/10.1016/j.cej.2017.11.088
Received 23 September 2017; Received in revised form 14 November 2017; Accepted 17 November 2017
Available online 21 November 2017
1385-8947/ Published by Elsevier B.V.
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

anaerobic digestion (AD), AHP offers unique advantages including high was developed by observing the fluorescent green dye in a vessel to
reaction rate, high operational stability, and clean energy production elucidate the mixing characteristics [23]. The biofermentor can be di-
with hydrogen gas as the major biogas component [1]. Previous re- vided into two sections – active and inactive volumes, with active vo-
searchers have found that several engineering parameters (e.g. pH lume being formed as a cavern around the mixer. It is expected that the
[2,3], temperature [4,5] and substrate content [5,6]) directly affected size of the active volume would change with the mixing position in AHP
hydrogen production (e.g. high hydrogen yield and suppression of the biofermentors, leading to variable mass distribution, gas bubble dis-
methanogens at pH of 4.8–5.5 and temperature of 60 °C [2]). In fact, the persion, biogas production, substrate precipitation and liquid fermen-
mixing of anaerobic systems is critical for reaction rate and energy tation production. Previous studies had found that in the journal-
consumption [7]. Normally, sufficient mixing ensures organic sub- bearing system (a system similar to the AHP experimental design in this
strates efficiently distributed in anaerobic systems to react with biomass study) consisting of two eccentric cylinders, the annular region was
particles, leading to fast heat transfer, release of trapped bubbles, and filled with a viscous liquid and the cylinders were rotated alternately to
avoid biomass precipitation [8]. Slow mixing (e.g. under 160 rpm) has produce chaotic advection [24]. Furthermore, the role of diffusion and
been used to facilitate sludge granulation, although complete-mixing transient velocities was studied in the dispersal of passive scalars pro-
AD systems can cause a washout of granular sludge [9,10]. Traditional duced in a low Reynolds number journal-bearing flow [25].
aerobic/anaerobic systems have historically used centrally stirred The breakthrough of this study was to elucidate the impact of
mixing (henceforth called centric mixing) where the flow is essentially chaotic mixing on the enhancement of biogas production and organic
rotated uniformly. Although this mixing pattern is well adopted and removal in an AHP biofermentor, with the objective to explore the re-
easily designed and operated, it might not optimize the interaction lationships between the chaotic mixing theme and energy output of
between substrates and biomass particles, especially in AHP systems AHP process. There were four tasks in this study. First, a lab-scale AHP
where biogas production can complicate the flow field. As one of the biofermentor was set up to use glucose, a typical carbon source for
major energy consumption factors in anaerobic systems, it is critical to hydrogen production bacteria, and operated in different chaotic mixing
better understand the mixing strategy and develop new energy-saving conditions. Second, the important parameters in AHP such as gas pro-
and efficient mixing patterns. duction, liquid components and chemical oxygen demand (COD) re-
Mixing in fluids is accomplished slowly by molecular diffusion in moval efficiency were investigated under different mixing impeller
laminar flows [11], which is greatly accelerated by turbulence and is positions. Third, a numerical finite element model was developed to
associated with large velocity gradients, especially in the small scales explore the flow and particle motion in the AHP system and determine
and high shear rates. For these reasons, turbulence is not suitable for the relationship between mixing and energy output. Finally, the sig-
mixing in biological situations where the high shear rates could damage nificance of applying chaotic mixing for anaerobic biofermentors was
bacterial cells and tear the granular sludge. In other situations such as explored in the context of scientific and industrial perspectives.
highly viscous flows, the flow remains laminar. A novel and funda-
mental mixing process rooted in dynamical systems theory has been
proposed [12] and termed chaotic advection. It was theorized that 2. Materials and methods
chaotic mixing would significantly increase the exposed surface area.
Flow changes direction continuously under the chaotic mixing scheme, 2.1. Anaerobic hydrogen producing (AHP) biofermentor setup and
which is expected to greatly enhance organic removal by maximizing operation
contact between biomass particles and substrates. Even though turbu-
lent mixing is common in high Reynolds number flows where the large A BIOFLO 110 Fermentor (New Brunswick Scientific, NJ) (effective
amount of energy drives particles (e.g., bacterial cells, flocs) in random volume of 2 L, height: 25 cm, radius: 6.5 cm) was used as the AHP
directions and produces a homogeneous fluid, chaotic mixing can also system. The vessel was sealed for air impermeability. An agitator with
occur at low Reynolds numbers when time-periodic oscillations with an impeller was installed for mixing. The radius of the impeller was
fixed amplitudes are imposed [13]. Chaotic advection causes simple 2 cm and the width of the impeller was 1.75 cm with 6 blades uniformly
non-turbulent flows to exhibit complicated particle trajectories that distributed around the circumference (Fig. 1). Before the tests, 20.625 g
result in enhanced mixing, and has been applied to low Reynolds glucose was fully dissolved in 200 ml distilled water and added to the
number plasma mixing without damaging bacterial cells, to enhancing vessel as the carbon source, followed by filling with 2 L nutrient solu-
heat transfer, and to creating long-chain polymers. Our group has ap- tion containing 4.0 g NH4HCO3, 2.6 g NaH2PO4, 200 mg MgSO4·7H2O,
plied this concept to a variety of wastewater and groundwater appli- 20 mg KCl, 20 mg Na2MoO4·2H2O, 20 mg CaCl2·2H2O, 15 mg
cations and scales [14–17]. Furthermore, parallel-competitive reaction MnSO4·7H2O and 5.6 mg FeCl2. The chemical oxygen demand (COD) of
systems have been simulated in a chaotic flow by solving the differ- the mixed solution was 10 g/L. The vessel was then sparged with ni-
ential convection–diffusion-reaction and it was shown convincingly trogen gas (99.995% high purity, Airgas Co., CT) for 20 min to achieve
that besides the relative rates of reaction, the nature of the chaotic flow the anaerobic environment. Organic soil (22 g) collected from a
is a determining factor in the evolution of the aqueous concentrations greenhouse at the University of Connecticut was used as the inoculum
[18]. The groundwork was presented in a previous study [19] by sol- for anaerobic bacteria to obtain the biomass particle concentration of
ving numerically the aforementioned differential equations. An ex- 10 g/L in the AHP system.
cellent state-of-the-art review of the numerous applications and theo- Previous studies had found that the disturbance of shock loading
retical background of chaotic advection was presented this year [20]. was better absorbed by the digester at low speed mixing conditions
Kinetic mixing models have been studied to elucidate the effect of (80 rpm) than at high speed mixing conditions (200 rpm) [26], since the
mixing speed/intensity on mass dispersion, biogas production, sub- bacteria in the anaerobic digestion process are more likely to live in a
strate removal efficiency and mass viscosity. Idealized models demon- balanced micro-ecosystem [27], and the low speed mixing conditions
strating chaotic advection include the point vortex model, tendril-whorl helped to maintain this good process in digestors. Furthermore, the
flow, the pulsed source-sink system, and the three-vortex system. Until energy dissipated by the system at low mixing was only 1/16 as the
now, mixing in AD systems has been limited to centrally located im- high mixing condition [26]. In this study, the mixing rate of the im-
pellers for the convenience of system configuration and operation [21]. peller was set at 50 rpm (the lowest mixing setup for the BIOFLO bio-
There have been only a few studies of chaotic mixing (off center) on fermentor). The temperature of the AHP biofermentor was kept at 38 °C
mass distribution and reaction [22] and no chaotic mixing has been by using a heating pad wrapped around the vessel.
applied in the field of AHP. In fact, mass distributions in AHP systems
would vary with the mixing position. A flow visualization technique

1384
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

Fig. 1. The mixing positions in the AHP biofermentor (a)


and the impeller (b).

2.2. Mixing positions in the AHP biofermentor test kit (HACH, CO) with a spectrophotometer (DR 2800, HACH, CO).
The pH was on line monitored using a pH meter connected to the AHP
After the glucose, nutrient solution and soil were added into the biofermentor. The constituents of the liquid samples were analyzed
AHP biofermentor, the liquid level height was 19.5 cm. Three mixing using a gas chromatograph (GC) (Agilent 6950, CA) equipped with an
heights (5, 10.5 and 16 cm from the bottom) and three mixing eccen- injector (7683B series) and flame ionization detector. The biogas
tricities (0, 2, 4 cm from the center of the AHP biofermentor) were samples collected from the AHP biofermentor using a 200 μl gas tight
examined individually to determine the effects of eccentrically stirred syringe (Fisher Scientific, MA) were analyzed with a GC (Agilent 6950,
(henceforth called eccentric) mixing to the AHP process (Fig. 1). The CA) equipped with a packed column and thermal conductivity detector.
matrix of the mixing height and eccentricity is listed in Table 1. All nine All the liquid and biogas samples were measured in duplicate.
setups were examined in triplicate. Statistical analyses were conducted as follows: ANOVA was used to
determine the significance of the effects the height and eccentricity
have on COD removal and biogas production. The T-test was used to
2.3. Sample collection from the AHP biofermentor during batch-mode tests
quantify the significance of the slope in the relationships between the
chaotic mixing coefficient (Lyapunov exponent λ) and AHP perfor-
Based on three mixing heights and three mixing eccentricities, a
mance.
total of nine groups of experiments were conducted (Table 1) in the
batch-mode individually for 45 h until no more gas was produced in the
headspace and no COD or pH value changed in the liquid phase. The 2.5. Chaotic mixing modeling and Lyapunov exponent
liquid sample (volume: 10 mL) was collected every 5 h using a sampling
pump connected with the AHP biofermentor, and filtered using a filter COMSOL Multiphysics (COMSOL, Inc., Burlington, MA, USA), a fi-
member (pore size: 4.5 μm, Millipore, MA) to remove the suspended nite element method based software package, was used to simulate the
solids and bacteria and then diluted 10 times using distilled water be- chaotic mixing under different impeller locations in the AHP bio-
fore analysis. The biogas sample was collected from the headspace fermentor. Specifically, sensitive dependence on initial conditions is an
using an airtight bottle (volume: 250 mL), The volume of the biogas important characteristic of chaotic motion, and close neighboring tra-
produced during each batch test was measured by a volumetric flow jectories are found to diverge. Furthermore, the rate of divergence is
meter (Cole Parmer, IL). exponential in chaotic systems, while regular trajectories are found to
separate only linearly in time. The number that distinguishes between
various types of orbits is termed as the Lyapunov exponent. A positive
2.4. Analytical method
Lyapunov exponent is the fundamental criterion for the system to be
considered chaotic.
The COD concentration was measured using a COD TNT plus vial
The Lyapunov exponent can be defined using the following ex-
ample. Assume two points in space are separated by a small distance ε0.
Table 1
Matrix of mixing height and eccentricity tested in the AHP biofermentor. Each of these points generates an orbit in space, so the separation of
two orbits is a function of time. The Lyapunov exponent measures the
Mixing position rpm Input Input pH Temperature (°C) dependence on initial conditions, termed as the separation ε (x0, y0, t).
COD (g/
The mean exponential rate of divergence of the two initially close orbits
Eccentricity (cm) Height (cm) L)
can be calculated using Eq. (1):
0 5 50 9983.33 7.54 38
2 5 50 9718.33 7.22 38 1 |ε (x 0,y0 ,t )|
λ = lim lim
4 5 50 10286.67 7.37 38 t →∞|ε0 → 0| t |ε0 | (1)
0 10.5 50 9523.33 7.55 38
2 10.5 50 9400 7.44 38 where λ is the Lyapunov exponent. For a discrete-time dynamical
4 10.5 50 9363.33 7.61 38 system exhibiting chaotic advection ‖εn ‖ ∼ ‖ε0 ‖e λn . This implies that a
0 16 50 9396.67 7.33 38
semi-logarithmic plot of the separation distance between a pair of
2 16 50 9926.67 7.41 38
4 16 50 9898 7.36 38 particles released very close to each other as a function of time should
appear linear; at least until boundary effects from a reactor (e.g., the

1385
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

edge, the bottom, or the impeller) become important. In this study, such
type of analysis was used to verify the chaotic status of the flow fields in
the AHP biofermentor.
In order to verify whether the AHP biofermentor was operated
under laminar flow conditions, the Reynolds number was calculated
(Eq. (2)):
R e = RiΩ(R o−Ri)/ ν (2)
where Ri and Ro, are the impeller and AHP biofermentor outer radii of
the apparatus, respectively, ν is the kinematic viscosity of the liquid and
Ω is the relative angular velocity with which the impeller rotates with
respect to the cylinder. For the AHP biofermentor operated at 38 °C, the
density of glucose solution was measured as 1007 ± 4.47 kg/m3, the
impeller and biofermentor outer radii of the apparatus, Ri and Ro, were
2 cm and 6.5 cm, respectively, and the kinematic viscosity of the liquid
was 1.032 × 10−6 m2/s resulting in a Reynolds number in the order of
700, which is indeed within the laminar flow regime.

3. Results and discussion

3.1. Effect of chaotic mixing conditions on liquid fermentation products

The main liquid fermentation products were acetic acid and butyric
acid under all mixing locations (Table 2), while the concentration of
ethanol, propionate and lactic acid were neglected (< 0.1 mM). No
methane was detected in the biogas content, indicating that only
acidogenesis and butyrate acetogenesis occurred in the AHP process.
Previous study had shown that hydrogen gas was produced at pH of
4.5–6.0, while acetoclastic methanogenesis occurred at pH > 5.7 with
acetic acid and butyric acid as the main liquid products [28]. In this
study, the pH value of the AHP biofermentor dropped rapidly from
∼7.4 to ∼4.8 within 5 h of the batch tests, and the biogas production
started after 6 h, indicating that the growth of methanogens was in-
hibited at low pH.
With glucose as the model substrate, acidogenesis and acetogenesis
started with the conversion of glucose to pyruvate through the glyco-
lytic pathway. Acetic acid, butyric acid or ethanol were produced (Eqs.
Fig. 2. The variation of COD removal efficiency with HAc ratio, mixing height and mixing
(3) and (4) [29]), and highly associated with hydrogen production
eccentricity in the AHP biofermentor (a: HAc ratio; b: mixing height; c: mixing eccen-
(4 mol H2/mole glucose for producing acetic acid and 2 mol H2/g glu- tricity).
cose for producing butyric acid).
C6 H12 O6 + 2H2 O→ 2CH3 COOH + 2CO2 + 4H2 (3) y z y
C x Hy Oz + ( x+ − )O2 → xCO2 + H2 O
4 2 2 (8)
C6 H12 O6 → CH3 CH2 CH2 COOH + 2CO2 + 2H2 (4)
Two other fermentation pathways can also be conducted by According to the mass balance equation of the COD of carbohydrate
Clostridium articum and Clostridium barkeri, leading to the production of (Eq. (8)), the COD concentrations of 1 mol/L acetic acid and butyric
propionate, ethanol, lactic acid and carbon dioxide (Eqs. (5)–(7), [29]). acid are 2 mol/L and 5 mol/L (Eq. (8)), respectively. Thereby, more
COD was generated through butyrate production process than acetate
C6 H12 O6 + 2H2 → 2CH3 CH2 COOH + 2H2 O (5) production process (5 mol/L COD comparing to 4 mol/L COD with
C6 H12 O6 → 2CH3 CH2 OH + 2CO2 (6) 1 mol/L glucose, Eqs. (3)–(5)). There was a positive relationship be-
tween the ratio of acetic acid and COD removal efficiency (Table 2,
C6 H12 O6 → 2CH3 CHOHCOOH + 2CO2 (7) Fig. 2a), demonstrating that higher ratio of acetic acid led to lower COD

Table 2
The liquid products and biogas products of the AHP biofermentor under nine mixing conditions.

Mixing position Input COD (g/L) Output COD removal efficiency (%)

Eccentricity (cm) Height (cm) HAc (mM) HBu (mg/L) Glucose (Mm) HAc fraction (%) COD (mg/L)

0 5 9983.33 19.8 28.4 5.5 36.87 7038.33 29.50


2 5 9718.33 20.1 27.5 3.5 39.33 6618.33 31.90
4 5 10286.67 21.5 28.4 3.9 39.96 6879.33 33.12
0 10.5 9523.33 20.5 26.4 4.7 39.73 6706.67 29.58
2 10.5 9400 20.5 26.1 2.7 41.58 6193.33 34.11
4 10.5 9363.33 21.5 26 1.2 44.15 5783.33 38.23
0 16 9396.67 20.1 26.5 3.7 39.96 6335 32.58
2 16 9926.67 22.3 27.1 2.4 43.05 6311.67 36.42
4 16 9898 23.9 25.9 0.4 47.61 5317.83 46.27

1386
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

concentration of effluent. Table 3


Theoretically, if all glucose in the influent (∼10,000 mg/L) were Hydrogen production in the AHP biofermentor under nine mixing conditions.
converted into acetic acid, the final COD concentration would be
Mixing position Input COD Output (gas)
∼6666.67 mg/L and the COD removal efficiency was ∼33.33%. (mg/L)
However, half of the COD removal efficiencies among nine mixing lo- Eccentricity (cm) Height (cm) H2 (mM) SHY (mole H2/
cations tested were higher than 33.33% (Table 2). Average COD re- mole glucose)
moval efficiency in this study (34.62%, Table 2) was slightly higher
0 5 9983.33 104.5 2.01
than previous studies (30% at OLR of 10 COD/L·d [30]), which was 2 5 9718.33 102.8 2.03
caused by the lower OLR rate in this study (4.6–5 g COD/L·d). COD 4 5 10286.67 107.9 2.01
removal efficiency had been found to increase as the organic loading 0 10.5 9523.33 103.5 2.09
rate (OLR, g/L·d) decreased in the AHP process [31]. In fact, the results 2 10.5 9400 103 2.10
4 10.5 9363.33 103.4 2.12
in this study corresponded well to the reported COD removal efficiency 0 16 9396.67 101.9 2.08
at similar OLR (36% at OLR of 4.7 COD g/L·d [32]). Furthermore, since 2 16 9926.67 109.6 2.12
acetic acid is the substrate for hydrogen production (Eq. (9)) [33], the 4 16 9898 110 2.13
consumption of acetic acid in AHP can also lower the COD concentra-
tion in the AHP effluent.
CH3 COOH + 2H2 O→ 2CO2 + 4H2 (9)
Mixing height and mixing eccentricity affected the COD removal
efficiency and the acetic acid ratios in liquid products. The maximum
COD removal efficiency (46.26%) was achieved at the chaotic mixing
position of α = 4 cm and h = 16 cm, while the minimum COD removal
efficiency (29.52%) occurred at the symmetric mixing position of α = 0
and h = 5 cm (Table 2). Previous study had found that in laminar fluid
flow, the mixture quality in biofermentors was severely degraded due to
the presence of isolated mixing regions at the symmetric scenario (the
impeller in the center) [34], since the structure of the global unstable
manifolds would never fill the entire flow domain [35]. In this study,
insufficient mixing under the symmetric mixing (the impellers in the
center of the AHP biofermentor) reduced the contact of the biomass
particles with glucose, and resulted in lower glucose degradation rate
and lower COD removal efficiency (Table 2, Fig. 2b and c).
At the chaotic mixing conditions (eccentricities of 2 cm and 4 cm),
the COD removal efficiencies increased with the mixing heights
(Fig. 2b), while no pattern was found at symmetric mixing conditions
(R2 < 0.8 at eccentricities of 0 cm, Fig. 2b). The possible reason was
that stronger chaotic movements of the fluid particles were formed
around the impeller, and more eddies and vortices were formed at
higher mixing positions [22]. In contrast, the chaotic movements in the
symmetric mixing were relatively weaker and the particles have
minimal interaction with each other [36], so that the COD removal in
the symmetric mixing conditions did not change substantially with the
mixing height. Furthermore, the COD removal efficiencies increased
with the mixing eccentricities at all three mixing heights (5, 10.5 and
16 cm) (Fig. 2c), which clearly validated the chaotic mixing enhanced Fig. 3. The variation of SHY values with mixing height (a) and mixing eccentricity (b) in
biochemical reactions. A two-way ANOVA using the mean values of the the AHP biofermentor.
experiments was conducted and showed that COD removal efficiency
depended significantly on both of the mixing eccentricity (p value: certain amount of hydrogen gas could be consumed for propionate
.0486) and the height (p value: .0887). Previous studies had suggested production (Eq. (5), although it did not occur much in this study). The
that the relative eccentricity (α/D, α: distance of the impeller to the specific hydrogen yields (SHY, mole H2/mole glucose) was much lower
center of the biofermentor, D: diameter of the biofermentor) of the than the stoichiometric hydrogen yield (Table 3). The maximum gas
impeller should be between 24% and 48% to obtain the best mixing production per mole glucose (2.13 mol H2/mole glucose) was slightly
quality [22]. In this study, the mixing eccentricity of 4 cm was 30.77%, lower than previous studies (2.04–2.72 mol H2/mole glucose [30],
well in the range previously reported. 2.32 mol H2/mole glucose [37]), which may be attributed to the lack of
pH control and pre-heat shocked soil inocula in this study, since the
3.2. Effect of chaotic mixing conditions on biogas production heated shocked soil was found to selectively favor the growth of
acidogenic bacteria [38].
Hydrogen production in the AHP biofermentor was lower than the Mixing height and mixing eccentricities clearly affected hydrogen
theoretical hydrogen yield (4 mol for each mole of glucose, Eq. (3)). production. The maximum SHY (2.13 mol H2/mole glucose) was
Based on the glucose amount (∼20.6 g) in the feeding solution and the achieved at the chaotic mixing position of eccentricity α = 4 cm and
reaction temperature (38 °C, 311.15 K), the maximum hydrogen pro- height h = 16 cm, while the minimum SHY (2.01 mol H2/mole glucose)
duction in this study should be 11.69 L. In the AHP process, the hy- occurred at the symmetric mixing position of α = 0 and h = 5 cm
drogen production was highly associated with OLR and hydrogen par- (Table 3). Like the COD removal efficiencies (Fig. 2b), the stronger
tial pressure. Glucose might be degraded through other pathways such chaotic movements in the eccentric mixing conditions (mixing eccen-
as butyrate acetogenesis, leading to a lower hydrogen gas production tricities of 2 cm and 4 cm) enhanced the SHY values with the mixing
(half of acidogenesis). Moreover, as the partial pressure increases,

1387
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

Fig. 4. Results from COMSOL simulation with impeller at


height of 16 cm for three mixing eccentricities in 3D view
and 2D bird’s view. a) for α = 0 cm, b) for α = 2 cm, c) for
α = 4 cm. The color of the slice plot shows the velocity
magnitude (m/s) and the red lines show the streamlines
associated with this flow simulation.

heights, while no pattern was found at the centric mixing conditions observed on the inner wall of the biofermentor (Fig. 1Sb). Previous
(Fig. 3a). The experimental data of hydrogen production was confirmed studies had found that when the impeller was placed at a relative high
by the COD removal efficiencies, since the COD removal efficiency was position, the lower part of the biofermentor was not disturbed [38], and
positively correlated with hydrogen production (Eqs. (3) and (4)). the bubbles (gas produced in the biofermentor) would not be affected.
It should be noted that hydrogen production increased with the However, hydrogen production would be strongly affected in this study
eccentricity at heights of 10.5 and 16 cm, but not related with eccen- when the impeller was put at a lower position, since gas bubbles were
tricity at the height of 5 cm (Fig. 3b). The main reason was that the trapped in the liquid contents and not contributed to biogas production
boundary effects were strong at the height of 5 cm and affected the (as shown in Fig. 1Sb), and thus interfering with the correlation be-
overall flow patterns. Furthermore, based on the modeling results, the tween the SHY values and eccentricity. Like the COD removal effi-
flow pattern was faster around the impeller (as shown in Fig. 4a–c). At ciency, a two-way ANOVA using the mean values of the experiments
the higher mixing position (the height of the impellers: 16 cm), more was also conducted to study the dependence of SHY on the mixing
power was dissipated on the liquid surface due to the faster flow speed, eccentricity and height. SHY was found to depend significantly only on
which helped releasing the trapped bubbles from the liquid phase to the the mixing height (p value: .0026) and not depend significantly on the
headspace level (Fig. 1Sa). In contrast, at the low mixing position (the mixing eccentricity (p value: .165). But this result could easily change
height of the impellers: 5 cm), the driving force of releasing the trapped with the addition of few extra data points.
bubbles from the liquid phase was weak, and thus more bubbles were

1388
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

3.3. Flow simulations of chaotic mixing in AHP

A numerical finite element model was developed using COMSOL


Multiphysics to simulate flow and particle motion in the AHP bio-
fermentor. The flow was simulated by solving the Navier-Stokes equa-
tions using the Arbitrary Lagrangian-Eulerian (ALE) technique [39]
(Eqs. (10) and (11)):
∂ρ ∂x
− . ∇ρ + ∇ . (ρu ) = 0
∂T ∂T (10)

∂u ∂x
ρ ⎛ − . ∇u ⎞ + ρ (u. ∇) u = ∇ . [−pI + τ ] + F
⎝ ∂T ∂T ⎠ (11)

where ∂ρ and ∂u are the derivatives of density and velocity, respec-


∂T ∂T
tively, with respect to mesh reference time (T), x is a function of angular
∂r
velocity and time, u = v + ∂t , in which v is the velocity vector in the
rotating coordinate system, and r is the position vector, p is pressure, τ
Fig. 5. Average particle separation distance analysis for impeller at height of 16 cm and at is shear stress, and F is external forces. Non-overlapped domains for the
different eccentricities 0 cm, 2 cm, and 4 cm away from the center. The Lyapunov ex- rotating and fixed parts were defined to simulate the flow field. A flow
ponent λ is calculated over the first 5–8 s of the process. continuity boundary condition was applied on the common boundaries
of fixed and rotating domains. A rotating interior wall boundary con-
Table 4
dition was applied on the walls of the impeller. In order to simulate a
Lyapunov exponent (1/s) calculated as a function of height and eccentricity. free surface condition on the top boundary of the AHP biofermentor, a
symmetry boundary condition was applied to avoid any shear stress
Lyapunov λ α = 0 cm α = 2 cm α = 4 cm development. Because the velocity of the fluid at a solid wall is zero
H = 5 cm 0.37511 0.22465 0.17411
relative to the wall, no-slip boundary conditions were assumed at the
H = 10.5 cm 0.22757 0.30097 0.42960 walls. The simulations employed an unstructured triangular mesh
H = 16 cm 0.18662 0.35799 0.43704 consisting of 357,303 domain elements, 20,758 boundary elements,
and 1717 edge elements.
To simplify the simulation, the velocity field was computed using
the frozen rotor assumption, which was a special case of a steady state
study. To study particles’ motion and their distances relative to each
other, the effect of different flow fields created by different impeller
mixing locations (as shown in Table 1) were explored. Specifically,
several pairs (10–12) of massless particles were released and scattered
all over the AHP biofermentor. The particles in each pair were initially
placed 1–3 mm apart and the separation distance of the particles in
individual pairs was calculated for the first 60 s of the simulation. To
ensure that the boundaries of fixed and rotating domains should not
impact the particles’ motion, the particle-continuity boundary condi-
tion was applied on the common boundaries of fixed and rotating do-
mains.
The streamlines and velocity field for three different locations of the
impeller were demonstrated (Fig. 4). The control experiment whereby
the impeller was located at the center was characterized by a more
consistent three-dimensional (3D) helical and two-dimensional (2D)
circular flow pattern (Fig. 4a). The streamlines from the control ex-
periment were not completely disordered and a circular pattern was
observed. In contrast, a more complex flow pattern was observed when
the impeller was eccentrically located at 2 cm and 4 cm off the center
(Fig. 4b and c), respectively. In this case the flow pattern was dis-
ordered, showing that the eccentricity of the impeller created a more
complex and possibly chaotic flow in comparison to the control ex-
periment (impeller in the center).
The particle separation distance for impeller at height of 16 cm and
at different eccentricities α (0 cm, 2 cm, and 4 cm) away from the center
was illustrated (Fig. 5). The Lyapunov exponent λ was calculated over
the first 5–8 s of the process, since after that point in time, boundary
effects were experienced due to the finite size of the biofermentor with
particles bouncing off or getting stuck at the walls. For instance, in the
control experiment, two of the particles got stuck to the biofermentor
wall after 7.6 s, so tracking their separation distance after that point
Fig. 6. Relationships between Lyapunov exponent λ and COD removal efficiency (a) and was impossible. The natural logarithm of particle separation distance
gas production (b).
with respect to time is plotted for the impeller located at 0 cm, 2 cm,
and 4 cm away from the center and at height of 16 cm (Fig. 5), re-
spectively. Based on the analysis, as the eccentricity increased, the

1389
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

Table 5
The comparison of energy input and output under centric mixing and eccentric mixing in AHP biofermentors.

Mixing positions Energy Lyapunov exponent COD removal Specific hydrogen Energy output in Energy output in Energy
inputa λ efficiency (%) yields H2b HAcc wastedd
Eccentricity Height

Centric mixing 1 0.228 29.13 1.873 0.156 0.126 0.092


Eccentric mixing (α = 2, h = 10.5) 1 0.301 34.07 1.945 0.162 0.144 0.056
Eccentric mixing (α = 4, h = 16) 1 0.437 46.24 1.951 0.163 0.158 0.002

a
Because the energy inputs were almost same at all mixing conditions, they were considered to be “1”.
b
Energy output in H2 and HAc was the fraction of H2 COD to total COD.
c
Energy output in HAc was the fraction of HAc COD to total COD.
d
Energy wasted was the fraction of unreacted glucose COD to total COD.

Lyapunov exponent became more positive, which was an indication of removal efficiency, high specific hydrogen yield (SHY, mole H2/mole
chaotic flow. glucose) and higher energy outputs (biogas phase and liquid product
The particle separation distance analysis was conducted for all nine phase) and less wasted energy (unreacted glucose). Under the laminar
impeller heights and eccentricity values (Table 4). The Lyapunov ex- fluid flow, the mixture quality increased at the eccentrically stirred
ponent λ was the largest for the middle and top impeller positions and mixing positions, allowing for a more sufficient contact between glu-
at the maximum eccentricity α. For those two height positions there cose and biomass particles, leading to more H2 and HAc output in
exist a monotonic increase in λ as a function of α. This, however, did higher λ value mixing conditions. Because the acetic acid fermentation
not hold true for the bottom impeller position (the height of 5 cm), the pathway has the highest theoretical hydrogen output (4 mol H2/mole
boundary effects were significant and affected the overall flow patterns. glucose, Eq. (3)), the energy output of acetic acid favored the energy
In addition, for a specific eccentricity α the height effect appeared to be output of hydrogen gas. The energy output of acetic acid increased
inconclusive. A two-way ANOVA showed that the effect of impeller more with λ than hydrogen gas (Table 5), since a portion of hydrogen
eccentricity was significant (p value: .0447), while the effect of impeller gas was consumed by other chemical reactions (Eq. (5)), and hydrogen
height was not significant (p value: .375) (Table 4). gas trapped in the liquid solution of the AHP biofermentor.
However, the eccentric mixing positions in this study were con-
ducted in a batch-mode, lab-scale biofermentor (volume: 2 L, radius of
3.4. Correlation of chaotic mixing simulation with AHP performance
biofermentor: 6.5 cm, radius of impeller: 2 cm). The impeller to bio-
fermentor ratio was about 0.3 in this study, which is larger than a ty-
The simulation results demonstrated well the correlation between
pical bioreactor in real-world wastewater treatment plants (ratio: 0.18
the mass distributions with the chaotic mixing positions in the AHP
[40]). Higher ratio of the impeller would make the boundary effects
biofermentor. At the height of 10.5 cm and 16 cm (minimal boundary
more significant and lower the efficiency of eccentric mixing. There-
effects), the Lyapunov exponent increased with the mixing eccentricity,
fore, better results are anticipated in a larger biofermentor. However,
representing a more eccentric mixing condition. Due to the boundary
the effectiveness of eccentric mixing for scale-up AHP systems needs to
effects from the bottom of the AHP biofermentor, the chaotic mixing at
be validated in future studies. Furthermore, the biomass particles (or-
the height of 5 cm exhibited the opposite behavior. The Lyapunov ex-
ganic soil) used in this study were greatly benefited from the chaotic
ponent also increased with the mixing height at eccentric mixing con-
mixing, while bacterial cells (colloids) and soluble reagents normally
ditions. These simulation results corresponded nicely to experimental
used in anaerobic systems might be less affected by the chaotic mixing.
results. At the height of 10.5 cm and 16 cm, the COD removal efficiency
and hydrogen production increased with the mixing eccentricity, while
it did not relate with the mixing eccentricity at the height of 5 cm. At 4. Conclusions
eccentricities of 2 cm and 4 cm, the COD removal efficiency and hy-
drogen production also increased with the mixing heights, but did not The use of eccentric mixing to enhance mass distribution and
relate with the mixing heights for the centric mixing condition. It was bioenergy outputs in AHP biofermentors was for the first time, com-
assumed that the flow mixing increased with the mixing eccentricity prehensively studied through both experimental and modeling metho-
and height, leading to more eccentric mixing conditions (higher dology. Eccentric mixing regimes were achieved under different mixing
Lyapunov exponent λ value) for sufficient mixing of biomass particles height and eccentricity. A numerical finite element model showed that
in the AHP biofermentor, so that the glucose and organic acids could be the Lyapunov exponent λ values (strength of chaotic mixing) increased
utilized efficiently. This assumption was validated by the strong posi- as a function of eccentricity at heights of 10.5 cm and 16 cm, and in-
tive correlations between λ (the simulated results) and COD removal creased as a function of height at eccentricities of 2 cm and 4 cm. Due to
efficiency and gas production (AHP experimental results, with the boundary effects, the Lyapunov exponent λ values were irregular at the
strong boundary effects at the height of 5 cm being excluded) (Fig. 6). A bottom position (the height of 5 cm). The experimental results of the
T-test on significance of slope was conducted and the result showed that batch mode AHP tests showed that COD removal efficiency and hy-
COD removal efficiency as a function of λ was significant (p value: drogen gas production had a strong positive correlation with the λ
.02 < p < .05) whereas SHY as a function of λ was only borderline values. Mass distribution became more sufficient and the AHP reactions
significant (p value: .1 < p < .2). proceeded more efficiently under chaotic mixing conditions. This lab-
The impact of eccentric mixing conditions on the AHP process was scale study clearly demonstrated that eccentrically stirred mixing ac-
explored in this study to gain a better understanding of the potential for celerated the interaction between biomass particles and organic sub-
enhanced bioenergy production in biofermentors. The correlation be- strates and ultimately enhanced the AHP processes.
tween the Lyapunov exponent λ (a proxy for the intensity of chaotic
mixing) and energy output (biogas and liquid products) in AHP pro-
cesses was established under different mixing positions (Table 5). At the Acknowledgement
same energy inputs (mixing rpm, mixing duration, and influent COD
concentration), a more eccentrically stirred mixing condition (asso- The graduate student (Yuankai Huang) is supported by the China
ciated with a larger Lyapunov exponent λ value) led to a higher COD Scholarship Council (CSC) Doctorate Program.

1390
Y. Huang et al. Chemical Engineering Journal 334 (2018) 1383–1391

Appendix A. Supplementary data Eng. J. 64 (1) (1996) 117–127.


[20] H. Aref, J.R. Blake, M. Budišić, et al., Frontiers of chaotic advection, Rev. Mod.
Phys. 89 (2) (2017) 025007.
Supplementary data associated with this article can be found, in the [21] C. Ratanatamskul, T. Saleart, Effects of sludge recirculation rate and mixing time on
online version, at http://dx.doi.org/10.1016/j.cej.2017.11.088. performance of a prototype single-stage anaerobic digester for conversion of food
wastes to biogas and energy recovery, Environ. Sci. Pollut. Res. 23 (8) (2016)
7092–7098.
References [22] H. Ameur, Effect of the shaft eccentricity and rotational direction on the mixing
characteristics in cylindrical tank reactors, Chin. J. Chem. Eng. 24 (12) (2016)
[1] A. Ghimire, L. Frunzo, F. Pirozzi, et al., A review on dark fermentative biohydrogen 1647–1654.
production from organic biomass: process parameters and use of by-products, Appl. [23] S.C. Low, N. Eshtiaghi, P. Slatter, et al., Mixing characteristics of sludge simulant in
Energy 144 (2015) 73–95. a model anaerobic digester, Bioprocess Biosyst. Eng. 39 (3) (2016) 473–483.
[2] J. Zhong, D.K. Stevens, C.L. Hansen, Optimization of anaerobic hydrogen and me- [24] Y. Fan, R.I. Tanner, N. Phan-Thien, A numerical study of viscoelastic effects in
thane production from dairy processing waste using a two-stage digestion in in- chaotic mixing between eccentric cylinders, J. Fluid Mech. 412 (2000) 197–225.
duced bed reactors (IBR), Int. J. Hydrogen Energy 40 (45) (2015) 15470–15476. [25] P. Dutta, R. Chevray, Inertial effects in chaotic mixing with diffusion, J. Fluid Mech.
[3] Y. Chen, J.J. Cheng, K.S. Creamer, Inhibition of anaerobic digestion process: a re- 285 (1995) 1–16.
view, Bioresour. Technol. 99 (10) (2008) 4044–4064. [26] X. Gomez, M.J. Cuetos, J. Cara, et al., Anaerobic co-digestion of primary sludge and
[4] Y. Xue, H. Liu, S. Chen, et al., Effects of thermal hydrolysis on organic matter so- the fruit and vegetable fraction of the municipal solid wastes: conditions for mixing
lubilization and anaerobic digestion of high solid sludge, Chem. Eng. J. 264 (2015) and evaluation of the organic loading rate, Renewable Energy 31 (12) (2006)
174–180. 2017–2024.
[5] F. Passos, J. Carretero, I. Ferrer, Comparing pretreatment methods for improving [27] G. Lettinga, The anaerobic treatment approach towards a more sustainable and
microalgae anaerobic digestion: thermal, hydrothermal, microwave and ultrasound, robust environmental protection, Water Sci. Technol. 52 (1–2) (2005) 1–11.
Chem. Eng. J. 279 (2015) 667–672. [28] C.H. Ting, D.J. Lee, Production of hydrogen and methane from wastewater sludge
[6] X. Fonoll, S. Astals, J. Dosta, et al., Anaerobic co-digestion of sewage sludge and using anaerobic fermentation, Int. J. Hydrogen Energy 32 (6) (2007) 677–682.
fruit wastes: evaluation of the transitory states when the co-substrate is changed, [29] C. Li, H.H.P. Fang, Fermentative hydrogen production from wastewater and solid
Chem. Eng. J. 262 (2015) 1268–1274. wastes by mixed cultures, Crit. Rev. Environ. Sci. Technol. 37 (1) (2007) 1–39.
[7] S. Ghanimeh, M. El Fadel, P. Saikaly, Mixing effect on thermophilic anaerobic di- [30] Y. Sharma, B. Li, Optimizing energy harvest in wastewater treatment by combining
gestion of source-sorted organic fraction of municipal solid waste, Bioresour. anaerobic hydrogen producing biofermentor (HPB) and microbial fuel cell (MFC),
Technol. 117 (2012) 63–71. Int. J. Hydrogen Energy 35 (8) (2010) 3789–3797.
[8] P. Kaparaju, I. Buendia, L. Ellegaard, et al., Effects of mixing on methane production [31] B. Rincón, R. Borja, J.M. González, et al., Influence of organic loading rate and
during thermophilic anaerobic digestion of manure: lab-scale and pilot-scale stu- hydraulic retention time on the performance, stability and microbial communities
dies, Bioresour. Technol. 99 (11) (2008) 4919–4928. of one-stage anaerobic digestion of two-phase olive mill solid residue, Biochem.
[9] Z.P. Zhang, K.Y. Show, J.H. Tay, et al., Rapid formation of hydrogen-producing Eng. J. 40 (2) (2008) 253–261.
granules in an anaerobic continuous stirred tank reactor induced by acid incuba- [32] S.V. Mohan, V.L. Babu, P.N. Sarma, Anaerobic biohydrogen production from dairy
tion, Biotechnol. Bioeng. 96 (6) (2007) 1040–1050. wastewater treatment in sequencing batch reactor (AnSBR): effect of organic
[10] Y. Hou, D. Peng, B. Wang, et al., Effects of stirring strategies on the sludge gran- loading rate, Enzyme Microb. Technol. 41 (4) (2007) 506–515.
ulation in anaerobic CSTR reactor, Desalin. Water Treat. 52 (34–36) (2014) [33] A.C. Basagiannis, X.E. Verykios, Catalytic steam reforming of acetic acid for hy-
6348–6355. drogen production, Int. J. Hydrogen Energy 32 (15) (2007) 3343–3355.
[11] A.D. Stroock, S.K.W. Dertinger, A. Ajdari, et al., Chaotic mixer for microchannels, [34] C. Galletti, E. Brunazzi, On the main flow features and instabilities in an unbaffled
Science 295 (5555) (2002) 647–651. vessel agitated with an eccentrically located impeller, Chem. Eng. Sci. 63 (18)
[12] H. Aref, Stirring by chaotic advection, J. Fluid Mech. 143 (1984) 1–21. (2008) 4494–4505.
[13] P. Hall, D.T. Papageorgiou, The onset of chaos in a class of Navier-Stokes solutions, [35] M.M. Alvarez, J.M. Zalc, T. Shinbrot, et al., Mechanisms of mixing and creation of
J. Fluid Mech. 393 (1999) 59–87. structure in laminar stirred tanks, AIChE J. 48 (10) (2002) 2135–2148.
[14] A.C. Bagtzoglou, N. Assaf-Anid, R. Chevray, Effect of chaotic mixing on enhanced [36] K.C. Ng, E.Y.K. Ng, Laminar mixing performances of baffling, shaft eccentricity and
biological growth and implications for wastewater treatment: a test case with unsteady mixing in a cylindrical vessel, Chem. Eng. Sci. 104 (2013) 960–974.
Saccharomyces cerevisiae, J. Hazard. Mater. 136 (1) (2006) 130–136. [37] Y. Sharma, B. Li, Optimizing hydrogen production from organic wastewater treat-
[15] A.C. Bagtzoglou, P.M. Oates, Chaotic advection and enhanced groundwater re- ment in batch reactors through experimental and kinetic analysis, Int. J. Hydrogen
mediation, J. Mater. Civil Eng. 19 (1) (2007) 75–83. Energy 34 (15) (2009) 6171–6180.
[16] A.C. Bagtzoglou, A. Novikov, Chaotic behavior and pollution dispersion char- [38] J. Wang, W. Wan, Comparison of different pretreatment methods for enriching
acteristics in engineered tidal embayments: a numerical investigation, JAWRA J. hydrogen-producing bacteria from digested sludge, Int. J. Hydrogen Energy 33 (12)
Am. Water Resour. Assoc. 43 (1) (2007) 207–219. (2008) 2934–2941.
[17] P. Zhang, S.L. DeVries, A. Dathe, et al., Enhanced mixing and plume containment in [39] F. Duarte, R. Gormaz, S. Natesan, Arbitrary Lagrangian-Eulerian method for Navier-
porous media under time-dependent oscillatory flow, Environ. Sci. Technol. 43 (16) Stokes equations with moving boundaries, Comput. Methods Appl. Mech. Eng. 193
(2009) 6283–6288. (45) (2004) 4819–4836.
[18] J.M. Zalc, F.J. Muzzio, Parallel-competitive reactions in a two-dimensional chaotic [40] I. Angelidaki, L. Ellegaard, B.K. Ahring, Applications of the Anaerobic Digestion
flow, Chem. Eng. Sci. 54 (8) (1999) 1053–1069. Process[M]//Biomethanation II, Springer, Berlin Heidelberg, 2003, pp. 1–33.
[19] F.J. Muzzio, M. Liu, Chemical reactions in chaotic flows, Chem. Eng. J. Biochem.

1391

You might also like