Regular article
Sparger design as key parameter to define shear conditions in pneumatic bioreactors

https://doi.org/10.1016/j.bej.2020.107529Get rights and content

Highlights

  • Airlift and stirred tank bioreactors exhibited similar shear environments.

  • Maximum shear rate was observed in the bottom region, around sparger holes.

  • Sparger design plays a key role in determining the maximum shear rate.

  • Maximum shear rate and shear frequency must be used for proper bioreactor comparison.

Abstract

The average shear rate (γ˙av) is a parameter used to characterize the shear environment in bioreactors, enabling comparison of the performances of different bioreactor models in terms of microorganism morphology and viability, and consequently bioproduct formation. Based on this approach, pneumatic bioreactors have been classified as low shear devices. However, the shear behavior cannot be generalized over a wide range of operating conditions, suggesting that the maximum shear rate (γ˙max) may be more suitable for the purpose of bioreactor performance comparison. Therefore, the aim of this work was to evaluate average and maximum shear rates in pneumatic bioreactors (bubble column and airlift), based on computational fluid dynamics (CFD) simulations. Concentric-duct and split airlift bioreactors exhibited higher γ˙av values, compared to the bubble column design, due to the nature of the liquid circulation patterns. The pneumatic bioreactors exhibited a significant (order of magnitude) difference between γ˙av (11.0–27.3 s−1) and γ˙max 4555 to 25,040 s−1), reflecting a non-uniform spatial distribution. The γ˙max values occurred close to the sparger holes and presented a linear relationship with gas injection velocity, which is dependent on the sparger geometry. In this way, sparger characteristics (number and diameter of sparger holes) defined γ˙max values in pneumatic bioreactors, showing that sparger should be properly designed in order to avoid excessive local shear rates.

Introduction

The selection of a bioreactor model for application in an aerobic bioprocess should consider not only the oxygen transfer capability, but also the shear environment, since biochemical processes are extremely sensitive to the shear intensity when fragile animal cells, plant cells, and filamentous microorganisms are used [1]. Excessive shear can cause cell damage, leading to viability loss and cell disruption or disintegration, so bioreactors should provide moderate or low shear environments, in order to avoid such effects [1,2].

The traditional parameter applied for evaluation of shear conditions is the average shear rate (γ˙av), which is extensively used for comparison of the performances of different bioreactor models. Several studies have evaluated this parameter, with the development of methodologies and correlations for bioreactor types including the bubble column [[3], [4], [5], [6], [7]], airlift [2,6,[8], [9], [10], [11]], and stirred tank [7,[12], [13], [14], [15], [16]]. Based on this approach, the effect of γ˙av on microbial growth, enzyme activity, and biomolecule production has been evaluated [17]. A general observation is that pneumatic bioreactors provide environments with lower shear, compared to stirred tank bioreactors [1,10,18,19].

However, there are exceptions to this general behavior. Studies have evaluated cellulase and xylanase production using different strains of Aspergillus sp., comparing the performance of pneumatic and stirred tank bioreactors [[20], [21], [22]]. Typically, higher enzyme yields were achieved using the pneumatic bioreactors (airlift and bubble column), which could be attributed to the low shear environments in pneumatic bioreactors. However, enzyme production by filamentous fungi is favored by a dispersed hyphal morphology. Consequently, higher enzyme production may be induced by the higher shear rate in pneumatic bioreactors, because exposure of the fungi to the shear environment leads to a predominantly dispersed morphology. On the other hand, when Aspergillus sp. grows with pelleted morphology, production of citric acid may increase [23].

The exceptions mentioned above regarding the effects of shear levels in pneumatic and stirred tank bioreactors have also been observed in other studies. Siedenberg et al. [24] cultivated Aspergillus awamori in stirred tank and airlift bioreactors, using wheat bran as carbon source and inductor of xylanase production. Analysis of fungal morphology showed pelleted growth in the conventional stirred tank bioreactor, while filamentous mycelium formation was observed in the airlift bioreactor, which could have been due to a higher shear environment in the latter device.

Cerri and Badino [17] correlated the production of clavulanic acid by Streptomyces clavuligerus with the average shear rate in 4 L stirred tank and 6 L concentric-duct airlift bioreactors. For the same initial oxygen transfer condition, the maximum value of the fermentation broth consistency index (Kmax), which is a rheological parameter related to the morphological structure of the mycelium, was lower for cultivation in the concentric-duct airlift bioreactor, compared to the stirred tank bioreactor. It was found that a higher shear rate was associated with greater fragmentation of the bacterial hyphae and a lower consistency index of the broth, indicating higher shear rates in the pneumatic bioreactor.

Jesus et al. [25] evaluated the hydrodynamics and oxygen transfer in stirred airlift and stirred tank bioreactors operated with xanthan gum solution, at specific air flow rates (φair) ranging from 0.5–1.5 vvm. At the same agitation speed (N), the values of the volumetric oxygen transfer coefficient (kLa) and average shear rate (γ˙av) were higher for the stirred airlift bioreactor, due to the additional liquid circulation flow pattern observed in this bioreactor. However, the shear level depended as much on the presence or absence of mechanical agitation as on the operating conditions (N and φair).

These previous findings question the established protocol for evaluation of the operation and performance of bioreactors, since they suggested that γ˙av may not be an ideal parameter to use for characterization of shear conditions in bioreactors. In fact, adoption of a maximum shear rate (γ˙max) would seem to be more suitable for the purpose of comparison among different bioreactor models, since it describes the worst condition to which a microorganism could be exposed. Depending on the γ˙max value, a single exposure to this condition could be sufficient to cause irreversible damage to the cells. The γ˙max represents a local value associated with a particular location inside the bioreactor. In stirred tank bioreactors, the region around the impeller exhibits the highest shear rate [26], while in airlift bioreactors, this worst shear condition is found in the bottom region, around the sparger holes [[27], [28], [29]]. Nonetheless, despite the effect that γ˙max has on cells, several other factors may affect the resistance of cells to shear, such as cell size and the mechanical resistance of the cell wall.

The shear conditions in different pneumatic bioreactors were evaluated by Thomasi et al. [6], who performed Streptomyces clavuligerus cultivations for clavulanic acid production in 5 L bubble column (BC), concentric-duct airlift (CDA), and split airlift (SA) bioreactors. Using the relation between the maximum value of the fermentation broth consistency index (Kmax) and the average shear rate (γ˙av), proposed by Cerri and Badino [17], Thomasi et al. [6]. obtained the following order for the average shear rate: γ˙av,CDA>γ˙av,SA>γ˙av,BC. The value of γ˙av appeared to be related to the capacity of the bioreactor to produce an internal circulation, since the lowest γ˙av values were found for the bubble column bioreactor, which had no well-defined liquid circulation pattern.

The evaluation of γ˙max can be performed using computational fluid dynamics (CFD) to solve conservative equations (continuity and momentum), with calculation of the flow fields and velocity profiles enabling estimation of local and average shear rates for different fluids and operating conditions. CFD has been extensively applied for evaluation of the performance of pneumatic bioreactors, mainly by analyzing variables such as liquid velocity [[27], [28], [29], [30], [31], [32], [33], [34], [35], [36], [37], [38], [39], [40], [41], [42]], global and local gas hold-up [[27], [28], [29], [30],32,33,[35], [36], [37], [38], [39], [40], [41],[43], [44], [45]], and volumetric oxygen transfer coefficient [28,31,38,43]. However, few studies using CFD to evaluate the shear conditions in pneumatic bioreactors are reported in the literature [[27], [28], [29],31,37,44]. Recently, Esperança et al. [46] showed that CFD was a suitable tool for estimation of the average shear rate (γ˙av) in pneumatic bioreactors, obtaining values for this parameter within an expected range. In the present work, the aim was to evaluate and compare the average shear rate (γ˙av) and maximum shear rate (γ˙max) in airlift and bubble column bioreactors, as well as to obtain a relation between γ˙max and the sparger characteristics of pneumatic bioreactors.

Section snippets

Bioreactors and experimental setup

In this study, CFD was employed to evaluate the shear rate in the following types of pneumatic bioreactor: bubble column (BC), split airlift (SA), and concentric-duct airlift (CDA).

The pneumatic bioreactors evaluated had two different working volumes: 5 and 10 L. Four different geometries of 10 L square cross section airlift bioreactor were simulated, comprising two concentric-duct airlift (CDA) models and two split airlift (SA) models, based on the work of Esperança et al. [47]. Three

Average shear rates in the pneumatic bioreactors

The average shear rate values obtained from the CFD simulations were calculated by adopting a volume averaging procedure for the spatial distribution of the shear rate throughout the bioreactor volume. Fig. 4 presents the γ˙av values for the 5 L pneumatic bioreactors. For the bubble column bioreactor, γ˙av ranged from 11–17.1 s−1, while for the split airlift and concentric-duct airlift bioreactors, this parameter varied from 12.7–27.3 s−1 and from 15.6–25.3 s−1, respectively. The loop (airlift)

Conclusions

Computational fluid dynamics (CFD) was used to evaluate the average shear rate (γ˙av) in three different pneumatic bioreactors (bubble column, concentric-duct airlift, and split airlift) operated with different fluids. At the same specific air flow rate, the airlift bioreactors (concentric-duct and split) exhibited higher γ˙av than the bubble column bioreactor, due to the liquid circulation pattern observed in the former devices.

The average shear rate (γ˙av) in the pneumatic bioreactors

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

CRediT authorship contribution statement

Mateus N. Esperança: Conceptualization, Methodology, Software, Visualization, Writing - original draft, Writing - review & editing. Caroline E. Mendes: Investigation. Guilherme Y. Rodriguez: Investigation, Software. Marcel O. Cerri: Conceptualization, Writing - review & editing. Rodrigo Béttega: Methodology, Writing - review & editing. Alberto C. Badino: Conceptualization, Supervision, Writing - review & editing, Funding acquisition.

Acknowledgments

The authors are grateful for the financial support provided by the Human Resources Program of the Brazilian National Agency of Petroleum, Natural Gas, and Biofuels (PRH/ANP-44), the National Council for Scientific and Technological Development (CNPq, grant 478472/2011-0), the São Paulo State Research Foundation (FAPESP, grants 2011/23807-1 and 2012/17756-8), and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES, Finance Code 001).

References (83)

  • D. Siedenberg et al.

    Production of xylanase by Aspergillus awamori on complex medium in stirred tank and airlift tower loop reactors

    J. Biotechnol.

    (1997)
  • S.S. de Jesus et al.

    Hydrodynamics and mass transfer in bubble column, conventional airlift, stirred airlift and stirred tank bioreactors, using viscous fluid: a comparative study

    Biochem. Eng. J.

    (2017)
  • R. Bannari et al.

    Mass transfer and shear in an airlift bioreactor: using a mathematical model to improve reactor design and performance

    Chem. Eng. Sci.

    (2011)
  • J.K. Huang et al.

    Development of a novel multi-column airlift photobioreactor with easy scalability by means of computational fluid dynamics simulations and experiments

    Bioresour. Technol.

    (2016)
  • P. Lestinsky et al.

    The effect of the draft tube geometry on mixing in a reactor with an internal circulation loop - A CFD simulation

    Chem. Eng. Process. Process Intensif.

    (2015)
  • S. Moradi et al.

    3 dimensional hydrodynamic analysis of concentric draft tube airlift reactors with different tube diameters

    Math. Comput. Modell.

    (2013)
  • R. Rzehak et al.

    Unified modeling of bubbly flows in pipes, bubble columns, and airlift columns

    Chem. Eng. Sci.

    (2017)
  • M. Simcik et al.

    CFD simulation and experimental measurement of gas holdup and liquid interstitial velocity in internal loop airlift reactor

    Chem. Eng. Sci.

    (2011)
  • T. Xu et al.

    CFD simulation of internal-loop airlift reactor using EMMS drag model

    Particuology

    (2015)
  • T. Zhang et al.

    A novel airlift reactor enhanced by funnel internals and hydrodynamics prediction by the CFD method

    Bioresour. Technol.

    (2012)
  • Q. Huang et al.

    CFD simulation of hydrodynamics and mass transfer in an internal airlift loop reactor using a steady two-fluid model

    Chem. Eng. Sci.

    (2010)
  • H.P. Luo et al.

    Local characteristics of hydrodynamics in draft tube airlift bioreactor

    Chem. Eng. Sci.

    (2008)
  • C.E. Mendes et al.

    Hydrodynamics of Newtonian and non-Newtonian liquids in internal-loop airlift reactors

    Biochem. Eng. J.

    (2016)
  • D. Siedenberg et al.

    Production of xylanase by Aspergillus awamori on synthetic medium in stirred tank and airlift tower loop reactors: the influence of stirrer speed and phosphate concentration

    J. Biotechnol.

    (1997)
  • M. Soos et al.

    Effect of shear rate on aggregate size and morphology investigated under turbulent conditions in stirred tank

    J. Colloid Interface Sci.

    (2008)
  • K. Wichterle et al.

    Shear rate on centrifugal pump impeller

    Chem. Eng. Sci.

    (1996)
  • Y. Chisti

    Animal-cell damage in sparged bioreactors

    Trends Biotechnol.

    (2000)
  • C.J. Liu et al.

    Effects of orifice orientation and gas-liquid flow pattern on initial bubble size

    Chin. J. Chem. Eng.

    (2013)
  • L. Luo et al.

    Hydrodynamics and mass transfer characteristics in an internal loop airlift reactor with different spargers

    Chem. Eng. J.

    (2011)
  • J.C. Merchuk et al.

    Studies of mixing in a concentric tube airlift bioreactor with different spargers

    Chem. Eng. Sci.

    (1998)
  • J.B. Snape et al.

    Liquid-phase properties and sparger design effects in an external-loop airlift reactor

    Chem. Eng. Sci.

    (1995)
  • R. Pishgar et al.

    Effect of aeration pattern and gas distribution during scale-up of bubble column reactor for aerobic granulation

    J. Environ. Chem. Eng.

    (2018)
  • M. Nishikawa et al.

    Heat-transfer in aerated tower filled with non-Newtonian liquid

    Ind. Eng. Chem. Process Des. Dev.

    (1977)
  • A. Schumpe et al.

    Viscous media in tower bioreactors - hydrodynamic characteristics and mass-transfer properties

    Bioprocess. Eng.

    (1987)
  • S.S. Thomasi et al.

    Average shear rate in three pneumatic bioreactors

    Bioprocess Biosyst. Eng.

    (2010)
  • J.A.S. Perez et al.

    Shear rate in stirred tank and bubble column bioreactors

    Chem. Eng. J.

    (2006)
  • L.K. Shi et al.

    Estimation of effective shear rate for aerated non-Newtonian liquids in airlift bioreactor

    Chem. Eng. Commun.

    (1990)
  • E.M. Grima et al.

    Characterization of shear rates in airlift bioreactors for animal cell culture

    J. Biotechnol.

    (1997)
  • A. Campesi et al.

    Determination of the average shear rate in a stirred and aerated tank bioreactor

    Bioprocess Biosyst. Eng.

    (2009)
  • M.M. Buffo et al.

    Influence of dual-impeller type and configuration on oxygen transfer, power consumption, and shear rate in a stirred tank bioreactor

    Biochem. Eng. J.

    (2016)
  • A.B. Metzner et al.

    Agitation of non-Newtonian fluids

    AlChE J.

    (1957)
  • Cited by (15)

    • A gas distributor capable of multiple injection directions to improve the gas–liquid dispersion performance in the airlift loop reactor

      2023, Biochemical Engineering Journal
      Citation Excerpt :

      Therefore, the structural type, and number and spacing of openings of gas distributor directly affect the fluid flow state and gas–liquid dispersion performance in the reactor, and the structural changes can affect the overall gas holdup and local gas fraction, which further affect the effect of wastewater treatment [6–9]. At present, the structural types of gas distributors mainly include single-orifice, sinter plate, perforated plate, porous plate, membrane, ring-type, and arm-type [10–13]. Most of these distributors are characterized by single structure, and the gas phase velocity is almost the axial velocity rising along the draft tube.

    • Hydrodynamics and mass transfer in spinner flasks: Implications for large scale cultured meat production

      2021, Biochemical Engineering Journal
      Citation Excerpt :

      However, the magnitude of the shear rate in the current design is much lower at about 100 1/s, as compared to 5,000–30,000 1/s in [37]. This is mainly due to the low gas exit velocity (volumetric gas flow rate divided by the surface area of the sintered sparger) of 0.06 m/s adopted by the current design, as opposed to 5-25 m/s considered by Esperança et al. [37]. It should be kept in mind that the method used here (i.e., Eq. (3)) tends to underestimate the contribution of turbulence to the shear stress.

    • Fungal bioprocessing of lignocellulosic materials for biorefinery

      2021, Recent Advancement in Microbial Biotechnology: Agricultural and Industrial Approach
    View all citing articles on Scopus
    View full text