Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023].

See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Received: 22 March 2018 | Revised: 24 July 2018 | Accepted: 30 July 2018

DOI: 10.1002/bit.26812

REVIEW

Progression of continuous downstream processing of


monoclonal antibodies: Current trends and challenges

Balaji Somasundaram1 | Kristina Pleitt1 | Evan Shave1,2 | Kym Baker2 |


1,3
Linda H. L. Lua

1
Australian Research Council Training Centre
for Biopharmaceutical Innovation, Australian Abstract
Institute for Bioengineering and Rapid advances in intensifying upstream processes for biologics production have left
Nanotechnology, The University of
Queensland, Brisbane, Queensland, Australia downstream processing as a bottleneck in the manufacturing scheme. Biomanufacturers
2
Patheon Biologics—a part of Thermo Fisher are pursuing continuous downstream process development to increase efficiency and
Scientific, Brisbane, Queensland, Australia
flexibility, reduce footprint and cost of goods, and improve product consistency and
3
Protein Expression Facility, The University of
Queensland, Brisbane, Queensland, Australia quality. Even after successful laboratory trials, the implementation of a continuous
process at manufacturing scale is not easy to achieve. This paper reviews specific
Correspondence
Linda H. L. Lua, Protein Expression Facility, challenges in converting each downstream unit operation to a continuous mode. Key
The University of Queensland, Brisbane, elements of developing practical strategies for overcoming these challenges are detailed.
Queensland 4072, Australia.
Email: l.lua@uq.edu.au These include equipment valve complexity, favorable column aspect ratio, protein‐A resin
selection, quantitative assessment of chromatogram peak size and shape, holistic process
Funding information
Australian Research Council, Grant/Award characterization approach, and a customized process economic evaluation. Overall, this
Number: IC160100027 study provides a comprehensive review of current trends and the path forward for
implementing continuous downstream processing at the manufacturing scale.

KEYWORDS
continuous manufacturing, continuous chromatography, process scale‐up and membrane
adsorbers

1 | INTRODUCTION 20 biopharma companies are expected to invest $109 billion in


research and development (R&D; Figure 1b,c) and the overall R&D
The global biopharmaceutical market is growing at a compound annual expenditure is expected to increase 2.8% each year to reach $182
rate of 10% since 2014 and is expected to reach approximately $390 billion by 2022 (EvaluatePharma, 2016). From a process development
billion by the end of 2019 (Ayturk & Marshall, 2016). Therapeutic perspective, most of the R&D is focused on intensifying and integrating
monoclonal antibodies (mAb) are a key class of biopharmaceutical bioprocess technologies to increase efficiency and flexibility in
products that has grown briskly in product approvals and sales, from biopharmaceutical manufacturing (Croughan, Konstantinov, & Cooney,
the time when the first mAb was commercialized (Ecker, Jones, & 2015; Estes & Longer, 2017). In this quest for faster and more efficient
®
Levine, 2015). As shown in Figure 1a, Humira , an antitumor necrosis processes, continuous manufacturing, designed to enhance efficiency,
factor‐α mAb manufactured by AbbVie (AbbVie Inc., North Chicago, IL), has emerged as a new modality in biomanufacturing.
achieved gross sales of $12.5 billion in 2014 (Udpa & Million, 2016) and This study elaborates on the current trends and potential risks in
increased by 28.8% in 2016 to reach $16 billion (Lindsley, 2017). The continuous downstream processing of mAbs with a primary focus on
biopharmaceutical market has become more dynamic with the recent continuous chromatography. We provide insights into the ways to
increase in the entry of biopharma companies into the biosimilars move from bench‐scale solutions to industrial implementation.
business (Udpa & Million, 2016). To maintain a competitive space in Furthermore, the article highlights the importance of a holistic
such a rapidly expanding market, in the next 5 years, the top approach to process development and characterization to address

Biotechnology and Bioengineering. 2018;115:2893–2907. wileyonlinelibrary.com/journal/bit © 2018 Wiley Periodicals, Inc. | 2893
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2894 | SOMASUNDARAM ET AL.

conditions optimization (Li, Lee, Zhou, Tressel, & Yang, 2006). High‐
throughput clone screening and selection using fluorescence‐acti-
vated cell sorting along with the implementation of miniaturized
bioreactors to optimize media, feeding regime and bioreactor
production has decreased the timeframe for cell culture process
development (Li et al., 2006). Rameez, Mostafa, Miller, and Shukla
(2014) have demonstrated the use of 10–15 ml advanced microscale
bioreactors (ambr™, TAP Biosystems, part of the Sartorius Stedim,
Greenville, Delaware) to optimize expression process parameters for
achieving optimum cell growth, viability, and product titer. The cell
culture profiles achieved in this microscale bioreactor study was
consistent with 3, 15, and 200 L stirred‐tank bioreactors.
For large‐scale production, biologics manufacturers either take a
fed‐batch approach for the ease of implementation (Lim, Sinclair,
Shevitz, & Bonham‐Carter, 2011), a perfusion process for increased
productivity (Kunert & Reinhart, 2016), or a combination of perfusion
and fed‐batch (W. C. Yang et al., 2014). Under constant supply of
fresh nutrients in a perfusion culture with enhanced cell densities,
mAb productivity reached 425 mg/L/day after 12 days, while over
the same period fed‐batch produced 55 mg/L/day (Langer & Rader,
2014). The evolution of continuous perfusion bioreactors and the
availability of established single‐use upstream technologies offer cost
and performance benefits to biologics manufacturers. A recent cost
of goods (CoGs) evaluation reported 30% savings can be achieved in
an integrated continuous process using disposable technologies over
a stainless steel batch process (Hummel et al., 2018). The upstream
process has also significantly advanced in the development of
process analytical tools for monitoring bioreactor performance,
meeting the requirements of the Food and Drug Administration
(FDA) for the implementation of continuous manufacturing. Various
in‐process, on‐line, and at‐line analytical tools used during upstream
mAb production have been reviewed in detail (Li, Vijayasankaran,
Shen, Kiss, & Amanullah, 2010). While the industry has witnessed
major developments in upstream processes, there has been little
progress in downstream processing. This has resulted in the
continuous accumulation of an increased amount of product (Clincke
F I G U R E 1 (a) 2014–2016 gross annual sales of products for et al., 2013) thereby making downstream processing a “bottleneck”
which biosimilars are in development (source: EvaluatePharma, in the manufacturing scheme (Gottschalk, 2008).
2016, Lindsley, 2017; Udpa & Million, 2016). (b) A 5‐year forecast
of R&D expenditure for the top 20 biopharma companies,
(e.g., 2016–2021; source: EvaluatePharma, 2016). (c) A 5‐year
3 | E V O L V I N G T R E N D S I N M ID S T R E A M
forecast of overall R&D expenditure for the biopharma companies
(source: EvaluatePharma, 2016) PROCESSING

Continuous centrifugation, depth filtration or tangential flow filtra-


the regulatory challenges related to introducing continuous down- tion (TFF) have been used as primary clarification techniques at
stream processing at manufacturing scale. manufacturing scale to remove the bulk of large particles, cell debris,
and whole cell impurities (Supporting Information Figure S1a;
Berthold & Kempken, 1994). Continuous centrifugation, operating
2 | EVOLVING TRENDS IN UPSTREAM at 10,000 to 12,000g with a short residence time, is not highly
PROC ESSIN G efficient at removing particles <1 µm (Yavorsky, Blanck, Lambalot, &
Brunkow, 2003). Moreover, at high cell densities a centrifugation
Upstream process development for mAb production typically process requires frequent desludging, leading to a decrease in
involves (a) clonal cell line selection and (b) media and bioreactor product recovery (Popova, Stonier, Pain, Titchener‐Hooker, & Farid,
T A B L E 1 Comparison of common clarification technologies (based on manufacturer data)

Feed stream
Acoustic wave Alternating tangential Continuous counter pretreatment with
Parameter separation flow filtration Continuous centrifugation flow centrifugation Depth filtration depth filtration
SOMASUNDARAM

Optimized total cell 20–50 20–146 3–13 6–150 10–30 >30


ET AL.

density at time of
clarification
(×106 cells/ml)a
Typical viability at time >60 >70 >40 >80 >40 >40
of clarification (%)a
Typical culture volume (L) 3–2,000 1–1,000 80–5,000b 0.1–6,000 0.1–2,000b 0.1–5,000b
Scalability Yes Yes Noc Semic Yes Yes
d
Continuous processing Yes Yes Yes Yesd No No
feasibility
References Collins and Clincke et al. Berthold and Kempken (1994), Masri (2017) Tomic et al. (2015)
Levison (2016) (2013) and Warikoo and Yavorsky et al. (2003)
et al. (2012)
a
Conditions at the time of harvest and operating parameters are process dependent and should be determined during the process development.
b
Technology is scalable beyond listed upper limit. Scale restrictions are based on infrastructure and when it becomes impractical from an economic standpoint.
c
While centrifugation has been used with culture volumes greater than 2,000 L, performance at small scale is not necessarily predictive of performance at large scale.
d
Operation lends itself to continuous processing, but auxiliary activities (e.g., assembly, disassembly, and cleaning) may reduce the benefits of continuous operation.
| 2895

10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2896 | SOMASUNDARAM ET AL.

2016). This limits the use of continuous centrifugation as respectively; Table 1). Among all available technologies, ATF and
a standalone unit operation for clarification and hence requires a continuous counter flow centrifugation can handle cell densities of
secondary clarification technique like depth filtration (Berthold & over 100 × 106 cells/ml, making them well suited for harvesting high
Kempken, 1994). Similarly, single stage clarification is not feasible cell density continuous perfusion processes (Table 1). The continuous
with depth filters or TFF. In a manufacturing set‐up, cleaning, counter flow centrifugation is a single use technology with 100 ml to
steaming, cooling, assembly, and disassembly of clarification hard- 6 L chamber capacity (Masri, 2017). Such a wide range of operational
ware is a significant rate limiting step (Yavorsky et al., 2003) and can flexibility is an advantage during continuous integrated process
potentially nullify the productivity advantages of an integrated development. Nevertheless, economic analysis on COGs and harvest
continuous process. Alternatively, clarification process with single‐ time for manufacturing scale bioreactors (e.g., 20,000 L production) is
use technology (SUT) using presterilized single‐use filters can critical for this new technology, particularly for less‐stable proteins.
increase process efficiency by eliminating cleaning, sterilization, and Among all midstream process technologies, ATF and TFF have
testing steps (Whitford, 2014). demonstrated capabilities of integrating with perfusion reactor and
Alternating tangential flow filtration (ATF) cell retention system Protein‐A (ProA) chromatography in a continuous mode, while
(Warikoo et al., 2012), acoustic wave separation (AWS) technology technologies like the AWS and depth filtration are used to clarify
(Collins & Levison, 2016), continuous counter flow centrifugation material from fed‐batch bioreactors to feed the continuous ProA
(Masri, 2017), and pretreatment of cell culture harvest with a chromatography step.
polycationic flocculating agent (e.g., pDADMAC; Burgstaller et al.,
2017; Tomic et al., 2015) are other clarification technologies
(Table 1). A 3–10‐fold reduction in secondary depth filter area, with a 4 | D O W N S T R E A M P L A T F O R M P RO C E S S
throughput of 15 L·m−2·h−1, was achieved by using AWS technology as
the primary clarification operation (Supporting Information Figure S1b; A mAb purification platform involving two to three chromatographic
Collins & Levison, 2016). Depth filtration of pDADMAC‐flocculated and filtration steps (Figure 2; Fahrner et al., 2001; Low, O’Leary, &
broths had a 4‐fold reduction in filter area when compared to traditional Pujar, 2007) typically accounts for 50–80% of the production costs of
two‐stage depth filtration (Burgstaller et al., 2017; Tomic et al., 2015). therapeutics (Farid, 2007; Gavara, Bibi, Sanchez, Grasselli, &
Using a tubular reactor, an online focus beam reflectance measurement Fernandez‐Lahore, 2015). The process suffers from the under-
and inline pressure sensor‐controlled valves, Burgstaller et al. (2017) utilization of the expensive ProA column to avoid breakthrough
demonstrated that depth filtration of flocculated cell culture could be (Mahajan, George, & Wolk, 2012) and biomanufacturing facilities
operated in a continuous manner. While flocculating agents are often trade‐off between production rate and operational costs by
beneficial in reducing depth filter area, demonstrating flocculant increasing the number of ProA cycles (typically 3–5 times) to reduce
clearance in the process intermediates using subsequent unit operations the column volume (Low et al., 2007; A. A. Shukla, Hubbard, Tressel,
is a challenge. Warikoo et al. (2012) have demonstrated the use of ATF Guhan, & Low, 2007). With the need to balance, the cost of raw
by directly loading the ATF clarified material onto the periodic counter‐ materials against production costs, organizational design (shift
current (PCC) system without additional clarification (Supporting management) and commercial demand, biomanufacturing facilities
Information Figure S1c). These technologies have shown to reduce or have come under pressure to improve the efficiency of their
bypass the need for secondary depth filtration, thereby decreasing downstream manufacturing process. For this purpose, integrated
operational footprint. continuous downstream processing is considered as a potential
Efficient clarification of cell culture harvest depends on a solution to reduce operational costs, increase efficiency and
combination of both total cell density and viability (Tomic et al., flexibility, streamline processes, reduce footprint, and improve
2015). AWS and flocculation technologies demonstrate capabilities of product consistency and quality (Bisschops & Brower, 2013; Gjoka,
handling high cell densities at a range of viabilities (25–50 × 106 cells/ Gantier, & Schofield, 2017; Hernandez, 2015; Jacquemart et al.,
ml at 60% viability and >30 × 106 cells/ml at 35–80% viability, 2016; Warikoo et al., 2012).

F I G U R E 2 A typical mAb purification


platform process involving three
chromatography unit operations
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SOMASUNDARAM ET AL. | 2897

5 | CONT IN UOUS DOW NSTREAM The first successful integration of a perfusion bioreactor and
PROC ESSIN G continuous protein capture using a GE Healthcare AKTA pcc 75 four‐
column PCC system (Uppsala, Sweden) was demonstrated in
Continuous chromatography utilizes multiple smaller columns to Warikoo et al. (2012). This study explored the operational flexibility
maximize the usage of resin’s binding capacity (Figure 3a). The use of and process train simplification of downstream process by removing
smaller columns provides an economic advantage by reducing intermediate hold steps. In 2014, Genzyme Corporation patented an
purification suite footprint, buffer and resin usage (Godawat et al., integrated and continuous process for manufacturing a therapeutic
2012). Increase in resin capacity coupled with shorter residence time protein drug substance (DS). The patented process included
can potentially improve productivity (g/Lresin/day; Hernandez, 2015; continuous ProA capture, viral inactivation, and continuous polishing
Mothes, Pezzini, & Schroeder‐Tittmann, 2016; Nestola et al., 2015; using two multicolumn chromatography systems. This integrated
Schaber et al., 2011). Moreover, continuous chromatography is continuous production process delivered DS with consistent product
suited for purification of less‐stable recombinant proteins, as the critical quality attributes (CQA; Konstantinov, Godawat, Warikoo, &
reduced process time helps ensure stability of the protein. The Jain, 2017). In the following year, an uninterrupted and fully
continuous chromatography process development has progressed automated end‐to‐end continuous purification of antibody DS with
tremendously in recent years due the availability of several a process productivity of >600 g/L resin/day was demonstrated by
commercial continuous chromatography systems (Table 2). Godawat, Konstantinov, Rohani, and Warikoo (2015). Taking a step

F I G U R E 3 Schematic representation of
downstream process intensification.
(a) Downstream processing involving
multicolumn continuous chromatography
within each unit operation.
(b) Downstream processing scheme
showing all unit operations performed
simultaneously and continuously.
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2898 | SOMASUNDARAM ET AL.

T A B L E 2 Comparison of commercial continuous chromatography system specification

Maximum operating Maximum number


Manufacturer Purification system Flow rate pressure (MPa) of columns Operating platform

Pall Life Science Cadence™ BioSMB PD Up to 70 or 1 16 Simulated moving bed


system 700 ml/mina
Cadence™ BioSMB 5–80 L/h 0.4 8
Process 80 system
Cadence™ BioSMB 30–350 L/h 0.4 8
Process 350 system
GE Healthcare AKTA™ pcc 75 1–75 ml/min 2 4 Periodic countercurrent
chromatography
BioProcess™ 1–60 L/h 2 4
PCC 4.7 mm
BioProcess™ PCC 6 mm 4–180 L/h 2 4
BioProcess™ PCC 10 mm 15–600 L/h 2 4
Novasepc BioSC® Lab Scale 1–30 ml/min 0.6 6 Sequential chromatography
®
BioSC Pilot Scale 30–1,500 ml/min 0.6 6
b
Semba Biosciences Semba ProPD™ Up to 70 or 100 1.86 8 Multicolumn continuous
or 300 ml/mina chromatography
ChromaCon Contichrom CUBE Up to 36 or 5 2 Periodic countercurrent
100 ml/mina chromatography
EcoPrime Twin GMP Up to 0.6 or 2.4 0.75 2
system or 9 L/mina
a
The flow rate depends on the pump selection.
b
Semba ProGMP™ system capable of processing 2,000 L is available (product specification to be released).
c
Customized commercial scale system.

further from the DS, the Novartis‐MIT Centre for Continuous Regardless of the economic benefits of operation in a continuous
Manufacturing is focusing on continuous manufacturing of the final mode, some limitations such as complexity in valve arrangements,
vialled drug product (Schaber et al., 2011). packing variability with multicolumn systems, operating robustness,
In another study, an integrated downstream bioprocess deliver- challenges in process characterization, validation and regulatory
ing purified mAb product was demonstrated using a 25 L fed‐batch concerns hold their implementation at larger manufacturing scales
bioreactor (Gjoka et al., 2017). An eight‐column ProA (5‐cm bed (Mothes et al., 2016; Stock, Bisschops, & Ransohoff, 2014).
height) process at a linear flow velocity of 360 cm/hr was used to
capture the mAbs. Following viral inactivation, the ProA eluate was
5.1 | Continuous protein A capture
further purified by an anion exchange membrane and mixed‐mode
cation exchange resin. The process was performed in an integrated ProA capture typically accounts for 25% (Costioli, Guillemot‐Potelle,
continuous mode over 30 cycles with 4.5 log reduction in host cell Mitchell‐Logean, & Broly, 2010) of total downstream cost. Supporting
proteins and 50% decrease in high molecular weight species (<2%). Information Figure S2 shows a breakdown of ProA capture cost,
Compared with a batch process, the integrated continuous 25‐L where consumables (e.g., resin) account for 67% of the cost. To offset
process showed superior economic advantage by utilizing 96% less the cost associated with ProA resin and to maximize the advantage of
ProA resin, 74% less anion exchange membrane, 97% less mixed‐ a continuous process, it is important to select a suitable ProA resin
mode cation exchange resin, and 44% less process buffer. This offers that has the ability to operate at a faster flow rate while retaining
a COGs advantage for high‐cost processes. high binding capacity and longer lifetime (Schaber et al., 2011).
As an alternative, an accelerated seamless antibody purification The first step in designing a ProA purification process is to select
approach was demonstrated by Sanofi (Paris, France; Mothes et al., an appropriate resin. Table 3 provides a comparison of characteristics
2016). In this process, all chromatography unit operations were for several commercially available ProA resins. In general, higher
performed in a continuous mode with only four buffers, eliminating dynamic binding capacity (DBC) is observed at a longer residence time
intermediate hold times (Figure 3b). As several batch processes are (RT). DBC is a measure of the amount of target protein the
carried out simultaneously, the time required to convert a batch chromatographic media can bind under respective flow conditions
process to continuous mode is minimal and the industry can define before a significant amount of target protein breakthrough (Do et al.,
and identify a unique batch more readily for regulatory purposes. 2008). Hence, DBC is as an important parameter reflecting mass
T A B L E 3 Comparison of protein A affinity media properties (based on manufacturer datasheets)

Maximum reported
Particle size DBC (mg human Residence
Manufacturer Resin name Base matrix (d50v; µm) IgG/mlresin) time (min) Ligand type Ligand attachment
SOMASUNDARAM

BioRad UNOsphere SUPrA™ Highly cross‐linked polymer 53–61 23–27 2 Recombinant protein A Epoxy
ET AL.

GE Healthcare MabSelect™ Highly cross‐linked agarose 85 30 2.4 Recombinant protein A Epoxy (single point)

MabSelect Xtra™ Highly cross‐linked agarose 75 40 2.4 Recombinant protein A Epoxy

MabSelect SuRe™ Highly cross‐linked agarose 85 35 2.4 Alkali‐stabilized protein Epoxy (single point)
A‐derived

MabSelect SuRe™ LX Highly cross‐linked agarose 85 60 6 Alkali‐stabilized protein Epoxy (single point)
A‐derived

MabSelect™ PrismA Highly cross‐linked agarose 60 80 6 Alkali‐stabilized Protein Epoxy (single point)
A‐derived

MabSelect SuRe™ pcc Highly cross‐linked agarose 50 60 2.4 Alkali‐stabilized protein Epoxy (single point)
A‐derived
Merck ProSep®‐vA High Capacity Controlled pore glass 74–125 >20 3–6 Native protein A NG
ProSep®‐vA Ultra Controlled pore glass 75–125 35 2.4 Native protein A (Multipoint)
®
ProSep Ultra Plus Controlled pore glass 60 >50 3–6 Recombinant structurally NG
conserved (native) protein A
Eshmuno A Hydrophilic polyvinyl ether 50 40–55 3–6 Proprietary ligand derived NG
from the C domain of S. aureus
Tosoh TOYOPEARL® AF‐rProtein Hydroxylated methacrylic 45 30 3 Alkali‐stabilized protein A (Multipoint)
A‐650F polymer

TOYOPEARL® AF‐rProtein A Hydroxylated methacrylic 45 65a 5a Alkali‐stabilized protein A (Multipoint)


HC‐650F polymer
Thermo Fisher POROS™ MabCapture™ Cross‐linked poly(styrene 45 ≥37 4 Recombinant native protein A Covalent
Scientific A divinylbenzene) immobilization
POROS™ MabCapture™ Cross‐linked poly(styrene 45 ≥37 4 Recombinant native protein A Covalent
A Select divinylbenzene) immobilization

Note. DBC: dynamic binding capacity; NG: not given .


a
At a 2‐min residence time, the binding capacity is 50 mg human IgG/mlresin.
| 2899

10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2900 | SOMASUNDARAM ET AL.

transfer limitations that may occur with the increase in flow rate. culture infectivity assays (Gombold et al., 2014; Johnson, Brown,
Apart from flow conditions, other major differences in the base matrix Lute, & Brorson, 2017). From an upstream perspective, continuous
type, particle size, ProA ligand density, modifications, and attachments perfusion mode bioreactors that run for months in comparison to
observed across different resins is also expected to influence the DBC fed‐batch bioreactors, requires more vigilance to ensure that novel
(Hahn et al., 2005; Lain, 2013). Higher mass transfer rates are infectious contaminations do not occur. Raw material sourcing, virus/
expected in resins with smaller particle diameter (Hahn, Schlegel, & mycoplasma filtration of cell culture media and new characterization
Jungbauer, 2003; Hahn et al., 2003). For instance, MabSelect Sure™ techniques have to be considered to reduce the risk of viral
pcc (GE Healthcare, Uppsala, Sweden) (50‐µm particle) and TOYO- contamination in the upstream stage. Failure to mitigate the risk of
PEARL® AF‐rProtein A HC‐650F (Tosoh Bioscience, Tokyo, Japan) viral contamination can lead to consequences such as facility
(45 µm particle) demonstrate high DBCs at shorter RT (60 mg/mlresin shutdown and decontamination of manufacturing facilities, for
at 2.4 min and 50 mg/mlresin at 2 min, respectively). The base beads of example, Genzyme’s production of Cerezyme® and Fabrazyme®
both resins are suitable for high mass transfer rates, but porous and was stopped (from 2009 to 2010 and 2012, respectively) due to the
rigid enough to accommodate faster flow rates with minimal pressure detection of a Vesivirus 2117 contamination at cell culture stage
drop across the chromatography bed. (Bestwick & Colton, 2009; Qiu et al., 2013).
The second important factor in selecting a ProA resin for Traditionally, in mAb downstream processes, viral clearance is
continuous processing is the resin reusability. A number of factors achieved by a combination of techniques including chromatographic
including operating conditions such as, CIP regime, feed stream, and separations, viral inactivation, and viral filtration methods (Connell‐
elution conditions that affect ligand stability, will impact resin Crowley et al., 2012; Farshid, Taffs, Scott, Asher, & Brorson, 2005;
lifetime (Jiang, Liu, Rubacha, & Shukla, 2009; Liu, Ma, Winter, & Miesegaes, Lute, & Brorson, 2010; Norling et al., 2005). In a batch
Bayer, 2010; A. S. Rathore & Sofer, 2012). As shown in Table 3 and mAb production process, a low pH virus inactivation step is a robust
Supporting Information Table S1, there are a variety of ligands and process that is routinely used for retrovirus inactivation (Bolton,
recommended CIP strategies based on ligand and base bead stability. Selvitelli, Iliescu, & Cecchini, 2015) for proteins, which are stable at a
Due to the common availability and demonstrated performance at transient low pH. However, in a continuous chromatography process
reducing bioburden, sodium hydroxide tends to be the preferred where the ProA eluate is collected continuously from different
method for cleaning (Biosciences, 2001; Burgoyne, Priest, Roche, & columns at different times, there is a risk of the eluate being held for
Vella, 1993; Girot, Moroux, Duteil, Nguyen, & Boschetti, 1990). As a a longer or shorter duration at low pH. mAbs are prone to
result, many resins have alkali‐stabilized ProA ligands. Even with aggregation when held at low pH for extended periods (Mazzer,
these modifications, repeated exposure to cell culture harvest and Perraud, Halley, O’Hara, & Bracewell, 2015). In an attempt to
sanitization solutions will degrade ProA resin, resulting in leached develop a continuous viral inactivation step, Klutz, Lobedann,
ProA (Carter‐Franklin, Victa, McDonald, & Fahrner, 2007) and Bramsiepe, and Schembecker (2016) evaluated a logarithmic reduc-
decreased binding capacity over time, leading to an increase in tion value approach and minimum residence time approach by using a
protein breakthrough (Hober, Nord, & Linhult, 2007). Leached ProA coiled flow inverter reactor. The study indicated that the minimum
ligand is a concern as studies have demonstrated it may cause residence time approach was more advantageous from a regulatory
immunogenic responses (Gómez et al., 2004). However, most of the perspective because ProA eluate leaves the reactor only after being
vendors have demonstrated that their resins will retain the majority held for the required inactivation time. In late 2017, Pall Corporation
of the binding capacity with the recommended CIP conditions for launched a fully automatic Cadence™ Virus Inactivation System (Pall
over 100 cycles and subsequent downstream unit operations are Corporation, Port Washington, New York) with the ability to perform
known to clear residual ProA leachates (Lain, 2013; Thillaivinayaga- continuous low pH virus inactivation. This system can be pro-
lingam, Reidy, Lindeberg, & Newcombe, 2012). Therefore, taking grammed to continuously collect ProA eluate, titrate the eluate to
DBC, RT, particle size, and resin lifetime into consideration, low pH, hold the eluate at low pH for a predefined time, and finally
MabSelect Sure™ PCC and TOYOPEARL® AF‐rProteinA HC‐650F titrate the treated eluate to high pH before discharge (Levison,
will be suitable resins for continuous processing requiring fast mass Schofield, Gantier, & Gjoka, 2017). Implementing this latest technol-
transfer, without compromising on binding capacity. ogy in mAb production process will require further studies on
process robustness, ability to integrate with different continuous
chromatography systems and process scale‐up.
5.2 | Continuous viral clearance
Solvent/detergent treatment and ultraviolet irradiation methods
Demonstrating virus safety is an important regulatory requirement are other alternative solutions for developing a continuous viral
for biologics expressed in human or animal cell lines (FDA, 1998). inactivation step that is complementary to continuous chromato-
Viral contamination of biologics, having the potential to impact graphy. In a recent review of implementing alternative methods for
patient safety, can occur due to the source of cell line or due to the assuring viral safety in a continuous downstream process, Johnson
introduction of adventitious agents during the manufacturing process et al. (2017) have highlighted the importance of evaluating virus
(Aranha, 2011). Currently, adventitious agents are tested via qPCR clearance in relation to resin age, particularly when the chromato-
assays, animal infection susceptibility, immune responses, or cell graphy resins are utilized to complete saturation during continuous
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SOMASUNDARAM ET AL. | 2901

downstream processing operations. In an end‐to‐end continuous of product to the AEX column and to improve separation of mAbs
process with sample flowing from one unit operation to another, from high molecular weight species.
spiking and sampling to titer virus input for the virus clearance In flowthrough mode purification, compared to the resin‐based
validation study can be challenging. To address this challenge, chromatography, membranes and monolith‐based chromatography is
Johnson et al. (2017) proposed a four‐valve system to ensure proper gaining more attention due to the use of faster flow rates due to
loading and mixing of viral load into viral clearance/inactivation unit convective flow, without compromising on adsorption capacity. For
operations and periodic sampling for qPCR to quantify genome copy example, Hughson, Cruz, Carvalho, and Castilho (2017) have
numbers and virus particle levels. The virus filtration unit operation, established the use of six different chromatography techniques,
the last purification step before the final processing, should be combining chromatography resins, membrane adsorbers, and mono-
carried out in a virus negative suite with facility controls that apply to liths for establishing a three‐step straight through purification
a batch process. However, integrating viral filtration operation strategy for purifying biopharmaceutical glycoproteins with minimal
carried out in an isolated suite with other unit operations for an or no sample adjustments in between unit operations. Although, this
end‐to‐end continuous process will require additional considerations. might be suitable for a lab scale process, it is critical to have sample
For example, the virus filter should be sized appropriately to handle adjustment steps and appropriate controls to measure product pH,
the higher load challenge in a continuous process or strategies on conductivity and concentration at the end of each unit operation at
using multiple virus filters in parallel can be developed by switching the manufacturing scale. The most commonly used membrane
the feed flow to the subsequent filter when flux decline is observed adsorbers for intermediate and polishing steps are anion exchange‐
in the first filter (Zydney, 2016). based membranes with positively charged quaternary ammonium
functional groups for adsorption of impurities (Zydney, 2016). The
newest generation of anion exchange membrane adsorbers contain-
5.3 | Continuous intermediate and polishing
ing covalently attached salt tolerant primary amine ion exchange
purification
ligands offer more evenly distributed binding capacity with enhanced
In the intermediate and polishing steps, most negatively charged pore accessibility and ligand density, across a conductivity range of
process related impurities such as DNA, some host cell proteins and 0–20 mS/cm. The ready scalability of membrane adsorbers in SUT
endotoxin (if present) bind to an anion exchange column, when format, eliminate the need for cleaning‐in‐place steps, making them
operating in product flow through mode (Knudsen, Fahrner, Xu, highly complementary to continuous ProA purification in increasing
Norling, & Blank, 2001; Li et al., 2010). The feed material for these productivity.
unit operations are often adjusted to the appropriate pH and
conductivity before loading (A. A. Shukla, Wolfe, Mostafa, & Norman,
5.4 | Continuous ultrafiltration/diafiltration
2017). During these adjustments mAbs can undergo chemical
degradation, aggregation, and modifications through oxidation, The final concentration and diafiltration of biologics DS is
reduction, deamidation, isomerization, and fragmentation, resulting regularly performed using TFF as a batch process. The biopharma
in a potential change to the amount of charge and size variants industry is targeting for high‐concentration protein formulations
(Khawli et al., 2010; Li et al., 2010; Weitzhandler et al., 1998). (up to 120 g/L) for subcutaneous administrations (Steele & Arias,
Designing a straight through chromatography step by connecting two 2014). For achieving such high concentrations, the process
polishing flow through steps, post ProA, can be considered as a solution is recirculated in the system several times, causing
potential solution to reduce the intermediate hold steps and the need additional shear and stress on the protein of interest. Single‐pass
for pH and conductivity adjustments. Klutz et al. (2015) demon- TFF (SP‐TFF) comes as an attractive alternative to traditional TFF
strated a straight through purification of neutralized ProA eluate system by using a unique flow path design and eliminating the
using a mixed mode and anion exchange chromatography in recirculation loop (Ayturk & Marshall, 2016). Recently, Rucker‐
flowthrough mode using a BioSMB system, with continuous sample Pezzini et al. (2018) addressed a major gap in the literature by
adjustment to reduce the ionic strength between the two‐chromato- demonstrating continuous diafiltration by integrating SP‐TFF into
graphy steps. Steinebach et al. (2017) demonstrated improved a fully continuous pilot scale production train to process 1 kg mAb
process performance with semicontinuous polishing step using a in 4 days. The study reported three identical stages of diafiltration
multicolumn countercurrent solvent gradient purification process. As in series to achieve 99.75% buffer exchange at a feed concentra-
an alternative to flowthrough intermediate and polishing steps, tion of 15 g/L and final product concentration of 148.5 g/L. The
Shamashkin, Godavarti, Iskra, and Coffman (2013) combined the current limitations on the availability of different size SP‐TFF
intermediate and polishing steps into a weak partition anion membrane is a bottleneck in using this technology for clinical and
exchange chromatography (WP‐AEX) and designed a semicontinuous commercial scale process (Rucker‐Pezzini et al., 2018). The impact
tandem process using ProA, WP‐AEX, and viral filtration. The precise of product concentration, formulation buffer type, pH, conductiv-
mobile phase requirements of WP‐AEX was achieved by high ity, excipient level, and process temperature on formulation offset
throughput optimization of ProA elution and neutralization buffers required to counter Donnan effect in SP‐TFF is another important
to achieve a stable ProA eluate pH. This prevented excessive binding knowledge gap in the literature.
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2902 | SOMASUNDARAM ET AL.

6 | MANU FAC TURING C HAL LEN GES column with 0.5‐mm diameter to conduct lab scale experiments to
derive mass balance, scale‐up and scheduling equations. In a recent
6.1 | Valve complexity study, Gjoka et al. (2017) showed the economic advantage of
continuous processing by using a 5‐cm ProA column at a flow velocity
Equipment manufacturers have incorporated additional valves, pH and
of 360 cm/hr. The column aspect ratios (diameter‐to‐bed height
conductivity probes, UV monitors, pressure gauges and control
ratio) operated in these laboratory scale studies usually allow good
hardware in a continuous chromatography system to ensure consis-
flow distribution unlike the process scale columns (Soriano, Titch-
tent product quality. In a GMP environment, these additional
ener‐Hooker, & Shamlou, 1997). Chromatography column scale‐up
components translate to increased complexity in regular operation
typically involves increasing the diameter of the column to process a
and process automation compared to a batch process (Jagschies, 2018;
higher quantity of load at a constant bed height and flow velocity
Stock et al., 2014). This complexity increases with the increase in
(Carta & Jungbauer, 2010). The diameters of manufacturing scale
number of columns. For example, in the Cadence™ BioSMB system,
columns can go above 1 m with corresponding aspect ratios >5
consists of 240 diaphragm valves in an integrated single use product‐
(Gerontas, Lan, Micheletti, & Titchener‐Hooker, 2015). Such changes
contact flow path used from 1 to 16 columns in a process development
in the column aspect ratio increases media compression and
scale or from one to eight columns in a manufacturing scale (Bisschops,
destabilizes the bed due to the loss of wall support and pressure
Frick, & Levison, 2016). Additional column positions can be added
drop (Stickel & Fotopoulos, 2001). This diminishing wall support
using this single‐use valve cassette containing all necessary inter-
directly impacts the chromatography column packing by lowering
connections during process development. However, such flexibility
critical flow velocity. In their paper, Kaltenbrunner, Diaz, Hu, and
may not be desired in a manufacturing set‐up as it can trigger a range
Shearer (2016) have characterized column dimensions to confirm
of documentation, such as additional calibration records, sanitization
that the continuous operation with shorter bed heights allows for an
records, qualification records, deviations, and corrective and preven-
increase in productivity and a decrease in buffer consumption at
tive action plans. On the contrary, GE Healthcare’s periodic counter
manufacturing scale. However, from the review of the literature, it
current technology with a minimum of three column configuration and
was identified that the favorable aspect ratios achieved in continuous
ChromaCon’s Eco Twin GMP system with two column configuration
chromatography has not been brought to light. In this review,
promises increase in productivity with minimum valve complications
chromatography column aspect ratios were calculated (Table 4) from
(Bisschops & Brower, 2013). From a manufacturing point of view, it
different column geometries for batch and continuous process
is important to design a process with equipment that offers a
reported by Bisschops et al. (2016). The calculated aspect ratio of
balance between maximizing productivity and operational simplicity.
continuous manufacturing process is 28–50% lesser, indicating that
Moreover, the incorporation of single use sensors for pH, conductivity,
the impact on critical flow velocity in a continuous process is 28–50%
product concentration, and pressure into a single use manifolds
lesser than the batch process. This analysis of aspect ratio should
can make the implementation of these complex systems more
install more confidence in biomanufacturers for continuous process
agreeable.
scale‐up.

6.2 | Column aspect ratio


6.3 | Column integrity testing
During continuous chromatography process development, shorter
columns are used at higher flow rates to maximize productivity. In a The column integrity testing before use, during processing, and after
proof of concept study, Pollock et al. (2013) used a 5‐cm bed height several runs is essential for the assurance of the quality performance

T A B L E 4 Comparison of column geometry of batch and continuous proA chromatography

Continuous clinical Continuous commercial Continuous commercial


Parameters Batch process manufacturing manufacturing manufacturing with high titer

Column diameter (cm) 80 20 30 30


Bed height (cm) 20 7 13 15
Aspect ratio 4.0 2.9 2.3 2.0
Reduction in aspect Not applicable 28 43 50
Ratio (%)
Resin volume (L) 100 17 39 50
Specific productivity (g/ 8.17 50 25 39
L/hr)
Column diameter
Aspect ratio = Bed height
(Aspect ratio of batch process − Aspect ratio of continuous process)
Reduction in aspect ratio = (Aspect ratio of batch process)
×100
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SOMASUNDARAM ET AL. | 2903

parameter W, the peak width at 50% peak height, is the most


commonly used measurement to characterize chromatograms
(Hinshaw, 2013). Peak width analysis has previously been used for
rapid protein kinetic characterization (Wang, Lewus, & Rathore,
2006), to characterize the separation of polar analytes and
chlorophenols on reversed phase liquid chromatography (Y. Yang,
Lamm, He, & Kondo, 2002) and analysis of ultraperformance liquid
chromatography results (Plumb et al., 2004). The 50% peak height
value is obtained by dividing the maximum UV absorbance recorded
at the peak apex by two. The width is determined by measuring the
volume between the leading and tailing edge at the calculated 50%
height. Variations in the calculated peak width is related to variations
in the number of theoretical plates (N) and HETP values using
F I G U R E 4 An example of start–peak–end–width (S–P–E–W)
equations described in Spangler and Collins (1975). A narrow peak
analysis. In this example, the conductivity gradient starts at 10 CV
width indicates an increase in N, while a wider peak is a result of a
and the elution peak collection beings at 20 CV. The peak start is
calculated as 10 CV from the start of the conductivity gradient to decrease in N. From a regulatory perspective, biomanufacturers are
start of peak collection. Similarly, peak apex is calculated as 20 CV expected to demonstrate consistency in column integrity (Siu, Chia,
from the start of the conductivity gradient to the elution peak apex Mok, & Pattnaik, 2014). Satisfying this regulatory expectation, the
(30 CV). Peak end is calculated as 30 CV from the start of the quantitative results obtained from the S–P–E–W analysis is a
conductivity gradient to the end of peak collection (40 CV). Peak
powerful tool to measure consistency in process chromatograms
width was calculated as 10 CV, the volume between 50% peak apex
on leading edge (25 CV) and 50% peak apex on tailing edge (35 CV) and column integrity across different columns used in a continuous
process, without increasing the process time.
The use of large‐scale prepacked columns for clinical and
of the chromatographic system. Moreover, the regulators expect the commercial manufacturing is becoming a growing trend in the
column to be suitably qualified before each use to ensure robust and biologics industry. The prepacked columns offer an economic
reproducible column performance (A. Rathore, Kennedy, Donnell, advantage in reducing column packing resources and associated
Bemberis, & Kaltenbrunner, 2003). The measure of height equivalent consumables. This advantage is more relevant for continuous
to theoretical plates (HETP) and the asymmetry (AS) is a conven- chromatography where multiple columns have to be packed. In a
tional method used to qualify the chromatography column (Endo, recent study conducted by Grier and Yakabu (2016), a prepacked
Nagamune, Katoh, & Yonemoto, 2000; Scharl et al., 2016). In ProA column (20 cm diameter and 10 cm bed height, axially packed)
continuous chromatography, incorporating column efficiency tests was compared to an axially packed stainless steel ProA column with
(HETP and AS measurement) after every cycle can increase the same dimensions. In‐process performance data, including product
process time and affect the scheduling of loading and nonloading yield and impurity clearance was comparable between the two
phases. As an alternative to this efficiency test, we propose to include columns. However, the study also reported discrepancies in the
the start–peak–end–width analysis, referred to as S–P–E–W, on the vendor and onsite column qualification data for the prepacked
process chromatograms as a measure of column integrity. S–P–E–W column. The difference in the results were attributed to continuous
is a simple quantitative characterization of elution peak size and media settling during and after shipment or system differences such
shape that is used to monitor variations in separation within a single as hold‐up volume or dissimilarities in the software used for analysis.
column over a period of time and across multiple columns operating Hence, developing a robust qualification process for using prepacked
in continuous mode. As shown in Figure 4, the peak start is calculated columns will be advantageous for implementing continuous process
as the volume (in column volumes) from the start of the conductivity at manufacturing scale.
gradient to the start of peak collection on the leading edge of the
peak. Similarly, peak apex and peak end values are calculated as the
6.4 | Process characterization
volume from the start of the conductivity gradient to the elution
peak apex and end of peak collection on the tailing edge respectively. The regulatory agencies strongly support continuous manufacturing,
In a continuous process, the S, P, and E measurements are useful data provided the manufactures develop a well understood robust process
to track retention shifts across different columns. A loss in ProA resin that can meet the product quality attributes under controlled
capacity is expected with the increase in the number of cycles (Jiang manufacturing steps (Nasr et al., 2017). However, the lack of
et al., 2009). In a continuous process, this loss in capacity occurs experience in getting regulatory approval for continuous bioproces-
earlier than the batch process due to overloading of the column sing creates an uncertainty in the industry (Stock et al., 2014). To
(Kaltenbrunner et al., 2016). Such degradation to stationary phase meet the regulatory levels of product quality assurance, the
resulting in retention shifts in a chromatography process (Parastar, continuous manufacturing process needs to be developed and
2015) is quantitatively identified using S, P, and E values. The final optimized under a set of critical process parameters (CPPs) enabling
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2904 | SOMASUNDARAM ET AL.

a rich knowledge‐based filing. In a traditional batch process, scale‐ process economics studies to suit their facility and optimize the
down model and characterization studies are conducted individually facility design to convert their existing sites, rather than assuming a
for each unit operation. However, while operating in a continuous fresh start (Jagschies, 2018).
mode, a design space incorporating multiple unit operations is
required for an increased operational flexibility. Such a multi-
dimensional risk analysis will assist biomanufacturers to achieve a 7 | CO NCL USION
holistic view of the process and identify CPPs interacting across all
unit operations, thereby defining a design space and control strategy In recent years, biopharma companies have evaluated continuous
across multiple process steps. To study such an extensive design operation of various downstream unit operations at laboratory scale to
space, biomanufacturers should consider establishing a scale‐down improve process efficiency while maintaining product quality. A major
model of the entire process train, rather than scaling down individual challenge in integrating continuous upstream and downstream processes
unit operations separately. Incorporating virus clearance studies into is in identifying and developing a suitable continuous midstream strategy.
such extensive scale‐down model to validate the clearance across Capital expense, COGs and scale‐down modeling capabilities will play a
each unit operation is another challenge for biomanufacturers. key role in developing a continuous midstream process. The favorable
column aspect ratios in continuous chromatography, the S–P–E–W
analysis detailed in this study, innovative resins and membranes with high
6.5 | Process economics
mass transfer and new chemistries, supported by rapid online analytics
Continuous processing offers the opportunity to increase manufac- will accelerate the transformation of this technology from bench‐top
turing flexibility, potentially improve product quality by operating in success to a manufacturing reality. A holistic approach to process
steady‐state, and reduce cost (Zydney, 2015). The allure of cost development and characterization will increase the process knowledge,
reduction has sparked numerous analyses to determine when and allowing the industry to make an informed strategic decision of either
where continuous processing can be most beneficial. A comparison developing an end‐to‐end continuous process or a fusion process that
study on the impact of batch versus continuous processing on COGs uses certain unit operations in continuous mode, while other unit
for an annual production of 200 kg of mAb showed a reduction in the operations are operated in batch mode at manufacturing scale.
downstream processing COGs by ~8€/g of mAb while the medium
requirements for continuous upstream processing increased the
upstream COGs by ~33€/g of mAb (Klutz, Holtmann, Lobedann, & AC KNO WL EDG M EN T

Schembecker, 2016). Therefore, a hybrid approach was recom- The authors acknowledge research funding under the Australian
mended. The advantages of a batch or a continuous or a hybrid Research Council Industrial Transformation Training Centre Program
process in terms of direct cost (e.g., raw materials, labor, QCQA (Project ID: IC160100027).
batch release), indirect cost (e.g., depreciation, facility‐dependent
overhead), manufacturing scale (i.e., clinical/commercial phase) may
ORCI D
vary based on the size of the company. For example, by using a
multiattribute decision‐making approach, Pollock, Coffman, Ho, and Balaji Somasundaram http://orcid.org/0000-0002-4555-2605
Farid (2017) have reported that a fully continuous and hybrid Linda H. L. Lua http://orcid.org/0000-0002-9756-7083
process can be more economically favorable in early phase clinical
production for small and medium sized companies. However,
continuous processing, particularly in upstream, was not as favorable R E F E R E N CE S
for late phase clinical production and larger companies. For all
company sizes, a complete batch process was the most beneficial Aranha, H. (2011). Virus safety of biopharmaceuticals. Contract Pharma, 13.
Ayturk, E., & Marshall, J. (2016). Using technology to overcome
during commercial manufacturing. Taking a slightly different
bioprocessing complexity: Advanced concentration and analytical
approach, Walther et al. (2015) assessed the economic benefits of technologies accelerate development and manufacture of mAbs,
an integrated continuous biomanufacturing platform for a hypothe- vaccines, and biosimilars. BioProcess International, 14(Suppl. 6), 20–24.
tical launch portfolio of mAbs and non‐mAbs using BioSolve process Berthold, W., & Kempken, R. (1994). Interaction of cell culture with
downstream purification: A case study. Cytotechnology, 15(1‐3), 229–242.
models. The study reported 57% reduction in capital expenses and
Bestwick, D., & Colton, R. (2009). Extractables and leachables from single‐
operation expenses compared fed‐batch production with single‐use use disposables. BioProcess International, 7(S1), 88–94.
technologies for the proposed product portfolio. Another considera- Biosciences, A. (2001). Use of sodium hydroxide for cleaning and
tion from a contract manufacturer perspective is that the reduction sanitizing chromatography media and systems. Process Chromatogr.
Bisschops, M., & Brower, M. (2013). The impact of continuous multi-
in process time for producing the target amount of protein for a
column chromatography on biomanufacturing efficiency. Pharmaceu-
project can result in the availability of time and resources for taking tical Bioprocessing, 1(4), 361–372.
more projects per year, resulting in an economic advantage. As the Bisschops, M., Frick, L., & Levison, P. (2016). Continuous chromatography
process economics results are derived from several general assump- is now possible for clinical manufacturing. BioProcess International,
14(6 s), 26–29.
tions, it will be appropriate for biomanufacturers to customize the
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SOMASUNDARAM ET AL. | 2905

Bolton, G. R., Selvitelli, K. R., Iliescu, I., & Cecchini, D. J. (2015). FDA (1998). Guidance for industry: Q5A viral safety evaluation of
Inactivation of viruses using novel protein A wash buffers. Biotechnol- biotechnological products derived from cell lines of human or animal
ogy Progress, 31(2), 406–413. https://doi.org/10.1002/btpr.2024 origin (ICH). Food and Drug Administration, Rockville, MD. (Online.).
Burgoyne, R. F., Priest, M. C., Roche, K. L., & Vella, G. (1993). Systematic http://www. fda. gov/downloads/RegulatoryInformation/Guidances/
development and validation of sanitization protocols for a chromato- UCM129101. pdf
graphic system designed for biotherapeutics purification. Journal of Gavara, P., Bibi, N., Sanchez, M., Grasselli, M., & Fernandez‐Lahore, M.
Pharmaceutical and Biomedical Analysis, 11(11), 1317–1325. https:// (2015). Chromatographic characterization and process performance
doi.org/10.1016/0731‐7085(93)80118‐K of column‐packed anion exchange fibrous adsorbents for high
Burgstaller, D., Krepper, W., Haas, J., Maszelin, M., Mohoric, J., Pajnic, K., throughput and high capacity bioseparations. Processes, 3(1),
… Satzer, P. (2017). Continuous cell flocculation for recombinant 204–221.
antibody harvesting. Journal of Chemical Technology and Biotechnology, Gerontas, S., Lan, T., Micheletti, M., & Titchener‐Hooker, N. J. (2015).
93, 1881–1890. Evaluation of a structural mechanics model to predict the effect of
Carta, G., & Jungbauer, A. (2010). Protein chromatography: process inserts in the bed support of chromatographic columns. Chemical
development and scale‐up. Weinheim, Germany: John Wiley & Sons Engineering Science, 129, 25–33.
Carter‐Franklin, J. N., Victa, C., McDonald, P., & Fahrner, R. (2007). Girot, P., Moroux, Y., Duteil, X. P., Nguyen, C., & Boschetti, E. (1990).
Fragments of protein A eluted during protein A affinity chromato- Composite affinity sorbents and their cleaning in place. Journal of
graphy. Journal of Chromatography A, 1163(1), 105–111. https://doi. Chromatography A, 510(C), 213–223. https://doi.org/10.1016/S0021‐
org/10.1016/j.chroma.2007.06.012 9673(01)93755‐0
Clincke, M.‐F., Mölleryd, C., Zhang, Y., Lindskog, E., Walsh, K., & Chotteau, Gjoka, X., Gantier, R., & Schofield, M. (2017). Transfer of a three step mAb
V. (2013). Very high density of CHO cells in perfusion by ATF or TFF chromatography process from batch to continuous: Optimizing
in WAVE Bioreactor™. Part I. Effect of the cell density on the process. productivity to minimize consumable requirements. Journal of
Biotechnology Progress, 29(3), 754–767. https://doi.org/10.1002/ Biotechnology, 242(Suppl. C), 11–18. https://doi.org/10.1016/j.
btpr.1704 jbiotec.2016.12.005
Collins, M., & Levison, P. (2016). Development of high performance Godawat, R., Brower, K., Jain, S., Konstantinov, K., Riske, F., & Warikoo, V.
integrated and disposable clarification solution for continuous (2012). Periodic counter‐current chromatography–design and opera-
bioprocessing. BioProcess International, 14, 30–33. tional considerations for integrated and continuous purification of
Connell‐Crowley, L., Nguyen, T., Bach, J., Chinniah, S., Bashiri, H., Gillespie, proteins. Biotechnology Journal, 7(12), 1496–1508.
R., … Vedantham, G. (2012). Cation exchange chromatography Godawat, R., Konstantinov, K., Rohani, M., & Warikoo, V. (2015). End‐to‐
provides effective retrovirus clearance for antibody purification end integrated fully continuous production of recombinant mono-
processes. Biotechnology and Bioengineering, 109(1), 157–165. clonal antibodies. Journal of Biotechnology, 213, 13–19. https://doi.org/
Costioli, M. D., Guillemot‐Potelle, C., Mitchell‐Logean, C., & Broly, H. 10.1016/j.jbiotec.2015.06.393
(2010). Cost of goods modeling and quality by design for developing Gombold, J., Karakasidis, S., Niksa, P., Podczasy, J., Neumann, K.,
cost‐effective processes. Biopharm International, 23(6). Richardson, J., … Sheets, R. L. (2014). Systematic evaluation of in
Croughan, M. S., Konstantinov, K. B., & Cooney, C. (2015). The future of vitro and in vivo adventitious virus assays for the detection of viral
industrial bioprocessing: Batch or continuous? Biotechnology and contamination of cell banks and biological products. Vaccine, 32(24),
Bioengineering, 112(4), 648–651. https://doi.org/10.1002/bit.25529 2916–2926.
Do, T., Ho, F., Heidecker, B., Witte, K., Chang, L., & Lerner, L. (2008). A Gottschalk, U. (2008). Bioseparation in antibody manufacturing: The good,
rapid method for determining dynamic binding capacity of resins for the bad and the ugly. Biotechnology Progress, 24(3), 496–503. https://
the purification of proteins. Protein Expression and Purification, 60(2), doi.org/10.1021/bp070452g
147–150. https://doi.org/10.1016/j.pep.2008.04.009 Grier, S., & Yakabu, S. (2016). Prepacked chromatography columns:
Ecker, D. M., Jones, S. D., & Levine, H. L. (2015). The therapeutic Evaluation for use in pilot and large‐scale bioprocessing. BioProcess
monoclonal antibody market. mAbs, 7(1), 9–14. https://doi.org/10. International, 14(4).
4161/19420862.2015.989042 Gómez, M. I., Lee, A., Reddy, B., Muir, A., Soong, G., Pitt, A., … Prince, A.
Endo, I., Nagamune, T., Katoh, S., & Yonemoto, T. (2000). Column qualification (2004). Staphylococcus aureus protein A induces airway epithelial
in process ion‐exchange chromatography. Paper presented at the Interna- inflammatory responses by activating TNFR1. Nature Medicine, 10(8),
tional Conference on Bioseparation Engineering: recovery and Recycle 842–848.
of Resources to Protect the Global Environment. Hahn, R., Bauerhansl, P., Shimahara, K., Wizniewski, C., Tscheliessnig, A., &
Estes, K. A., & Longer, E. (2017). Update on continuous bioprocessing: Jungbauer, A. (2005). Comparison of protein A affinity sorbents: II. Mass
From the industry’s perception to reality. Biopharm International, transfer properties. Journal of Chromatography A, 1093(1), 98–110.
30(6), 10–12. Hahn, R., Schlegel, R., & Jungbauer, A. (2003). Comparison of protein A
EvaluatePharma. (2016). World Preview 2016, Outlook to 2022. affinity sorbents. Journal of Chromatography B, 790(1), 35–51. https://
EvaluatePharma. Available online at: http://info. evaluategroup. doi.org/10.1016/S1570‐0232(03)00092‐8
com/rs/607‐YGS‐364/images/wp16. pdf (Accessed January 16, 2017). Hernandez, R. (2015). Continuous manufacturing: A changing processing
Fahrner, R. L., Knudsen, H. L., Basey, C. D., Galan, W., Feuerhelm, D., paradigm. Biopharm International, 28(4), 20–27.
Vanderlaan, M., & Blank, G. S. (2001). Industrial purification of Hinshaw, J. V. (2013). How do your peaks measure up? LC GC North
pharmaceutical antibodies: Development, operation, and validation of America, 31(10), 860+.
chromatography processes. Biotechnology and Genetic Engineering Hober, S., Nord, K., & Linhult, M. (2007). Protein A chromatography for
Reviews, 18(1), 301–327. antibody purification. Journal of Chromatography B, 848(1), 40–47.
Farid, S. S. (2007). Process economics of industrial monoclonal antibody Hughson, M. D., Cruz, T. A., Carvalho, R. J., & Castilho, L. R. (2017).
manufacture. Journal of Chromatography B, 848(1), 8–18. https://doi. Development of a 3‐step straight‐through purification strategy
org/10.1016/j.jchromb.2006.07.037 combining membrane adsorbers and resins. Biotechnology Progress.,
Farshid, M., Taffs, R., Scott, D., Asher, D., & Brorson, K. (2005). The 33(4), 931–940.
clearance of viruses and transmissible spongiform encephalopathy Hummel, J., Pagkaliwangan, M., Gjoka, X., Davidovits, T., Stock, R.,
agents from biologicals. Current Opinion in Biotechnology, 16(5), Ransohoff, T., … Schofield, M. (2018). Modeling the downstream
561–567. processing of monoclonal antibodies reveals cost advantages for
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2906 | SOMASUNDARAM ET AL.

continuous methods for a broad range of manufacturing scales. Lindsley, C. W. (2017). New 2016 data and statistics for global
Biotechnology Journal, 1700665. https://doi.org/10.1002/biot. pharmaceutical products and projections through 2017. ACS
201700665 Chemical Neuroscience, 8(8), 1635–1636. https://doi.org/10.1021/
Jacquemart, R., Vandersluis, M., Zhao, M., Sukhija, K., Sidhu, N., & Stout, J. acschemneuro.7b00253
(2016). A single‐use strategy to enable manufacturing of affordable Liu, H. F., Ma, J., Winter, C., & Bayer, R. (2010). Recovery and purification
biologics. Computational and Structural Biotechnology Journal, 14, 309– process development for monoclonal antibody production. mAbs, 2(5),
318. https://doi.org/10.1016/j.csbj.2016.06.007 480–499. https://doi.org/10.4161/mabs.2.5.12645
Jagschies, G. (2018). Chapter 28 ‐ c ontinuous capture of mAbs—Points to Low, D., O’Leary, R., & Pujar, N. S. (2007). Future of antibody purification.
consider and case studies, Biopharmaceutical Processing (527–556). Journal of Chromatography B, 848(1), 48–63. https://doi.org/10.1016/j.
London, United Kingdom: Elsevier. jchromb.2006.10.033
Jiang, C., Liu, J., Rubacha, M., & Shukla, A. A. (2009). A mechanistic study Mahajan, E., George, A., & Wolk, B. (2012). Improving affinity chromato-
of protein A chromatography resin lifetime. Journal of Chromatography graphy resin efficiency using semi‐continuous chromatography.
A, 1216(31), 5849–5855. Journal of Chromatography A, 1227, 154–162. https://doi.org/10.
Johnson, S. A., Brown, M. R., Lute, S. C., & Brorson, K. A. (2017). Adapting 1016/j.chroma.2011.12.106
viral safety assurance strategies to continuous processing of Masri, Q. A. R. F. (2017). Downstream processing for cell‐based therapies.
biological products. Biotechnology and Bioengineering, 114(1), 21–32. Biopharm International, 30(4).
https://doi.org/10.1002/bit.26060 Mazzer, A. R., Perraud, X., Halley, J., O’Hara, J., & Bracewell, D. G. (2015).
Kaltenbrunner, O., Diaz, L., Hu, X., & Shearer, M. (2016). Continuous bind‐ Protein A chromatography increases monoclonal antibody aggrega-
and‐elute protein A capture chromatography: Optimization under tion rate during subsequent low pH virus inactivation hold. Journal of
process scale column constraints and comparison to batch operation. Chromatography A, 1415, 83–90. https://doi.org/10.1016/j.chroma.
Biotechnology Progress, 32(4), 938–948. https://doi.org/10.1002/ 2015.08.068
btpr.2291 Miesegaes, G., Lute, S., & Brorson, K. (2010). Analysis of viral clearance
Khawli, L. A., Goswami, S., Hutchinson, R., Kwong, Z. W., Yang, J., Wang, X., unit operations for monoclonal antibodies. Biotechnology and Bioengi-
… Motchnik, P. (2010). Charge variants in IgG1: Isolation, character- neering, 106(2), 238–246.
ization, in vitro binding properties and pharmacokinetics in rats. mAbs, Mothes, B., Pezzini, J., & Schroeder‐Tittmann, K. (2016). Accelerated,
2(6), 613–624. https://doi.org/10.4161/mabs.2.6.13333 seamless antibody purification: Process intensification with continuous
Klutz, S., Holtmann, L., Lobedann, M., & Schembecker, G. (2016). Cost disposable technology, 11. bioprocessintl. com, May.
evaluation of antibody production processes in different operation Nasr, M. M., Krumme, M., Matsuda, Y., Trout, B. L., Badman, C., Mascia, S., …
modes. Chemical Engineering Science, 141, 63–74. Lee, S. L. (2017). Regulatory perspectives on continuous pharmaceutical
Klutz, S., Lobedann, M., Bramsiepe, C., & Schembecker, G. (2016). manufacturing: Moving from theory to practice: September 26‐27, 2016,
Continuous viral inactivation at low pH value in antibody manufactur- International Symposium on the Continuous Manufacturing of Pharma-
ing. Chemical Engineering and Processing: Process Intensification, 102 ceuticals. Journal of Pharmaceutical Sciences, 106(11), 3199–3206.
(Suppl. C), 88–101. https://doi.org/10.1016/j.cep.2016.01.002 Nestola, P., Peixoto, C., Silva, R. R. J. S., Alves, P. M., Mota, J. P. B., &
Klutz, S., Magnus, J., Lobedann, M., Schwan, P., Maiser, B., Niklas, J., … Carrondo, M. J. T. (2015). Improved virus purification processes for
Schembecker, G. (2015). Developing the biofacility of the future based vaccines and gene therapy. Biotechnology and Bioengineering, 112(5),
on continuous processing and single‐use technology. Journal of 843–857. https://doi.org/10.1002/bit.25545
Biotechnology, 213, 120–130. Norling, L., Lute, S., Emery, R., Khuu, W., Voisard, M., Xu, Y., … Brorson, K.
Knudsen, H. L., Fahrner, R. L., Xu, Y., Norling, L. A., & Blank, G. S. (2001). (2005). Impact of multiple re‐use of anion‐exchange chromatography
Membrane ion‐exchange chromatography for process‐scale antibody media on virus removal. Journal of Chromatography A, 1069(1), 79–89.
purification. Journal of Chromatography A, 907(1), 145–154. https:// Parastar, H. (2015). Multivariate curve resolution methods for qualitative
doi.org/10.1016/S0021‐9673(00)01041‐4 and quantitative analysis in analytical chemistry, Data handling in science
Konstantinov, K., Godawat, R., Warikoo, V., & Jain, S. (2017). Integrated and technology (29, 293–345). Oxford, United Kingdom: Elsevier.
continuous manufacturing of therapeutic protein drug substances. In: Plumb, R., Castro‐Perez, J., Granger, J., Beattie, I., Joncour, K., & Wright, A.
Google Patents. (2004). Ultra‐performance liquid chromatography coupled to quadru-
Kunert, R., & Reinhart, D. (2016). Advances in recombinant antibody pole‐orthogonal time‐of‐flight mass spectrometry. Rapid Communica-
manufacturing. Applied Microbiology and Biotechnology, 100, 3451–3461. tions in Mass Spectrometry, 18(19), 2331–2337.
https://doi.org/10.1007/s00253‐016‐7388‐9 Pollock, J., Bolton, G., Coffman, J., Ho, S. V., Bracewell, D. G., & Farid, S. S.
Lain, B. (2013). Protein A: The life of a disruptive technology. BioProcess (2013). Optimising the design and operation of semi‐continuous
International, 11(8). affinity chromatography for clinical and commercial manufacture.
Langer, E. (2014). Continuous bioprocessing and perfusion: Wider Journal of Chromatography A, 1284(Suppl. C), 17–27. https://doi.org/
adoption coming as bioprocessing matures. BioProcess Journal, 13(1), 10.1016/j.chroma.2013.01.082
43–49. https://doi.org/10.12665/J131.Langer Pollock, J., Coffman, J., Ho, S. V., & Farid, S. S. (2017). Integrated
Levison, P., Schofield, M., Gantier, R., & Gjoka, X. (2017). Implementation continuous bioprocessing: Economic, operational and environmental
of an end‐to‐end continuous bioprocessing platform using novel feasibility for clinical and commercial antibody manufacture. Biotech-
technologies. nology Progress., 33(4), 854–866.
Li, F., Lee, B., Zhou, J., Tressel, T., & Yang, X. (2006). Current therapeutic Popova, D., Stonier, A., Pain, D., Titchener‐Hooker, N. J., & Farid, S. S.
antibody production and process optimization. BioProcessing Journal, (2016). Integrated economic and experimental framework for screen-
5(4), 16–25. https://doi.org/10.12665/J54.LiZhou ing of primary recovery technologies for high cell density CHO
Li, F., Vijayasankaran, N., Shen, A., Kiss, R., & Amanullah, A. (2010). Cell cultures. Biotechnology Journal, 11(7), 899–909. https://doi.org/10.
culture processes for monoclonal antibody production. mAbs, 2(5), 1002/biot.201500336
466–477. https://doi.org/10.4161/mabs.2.5.12720 Qiu, Y., Jones, N., Busch, M., Pan, P., Keegan, J., Zhou, W., … Mattaliano, R.
Lim, J., Sinclair, A., Shevitz, J., & Bonham‐Carter, J. (2011). An economic J. (2013). Identification and quantitation of vesivirus 2117 particles in
comparison of three cell culture techniques. Biopharm International, bioreactor fluids from infected Chinese hamster ovary cell cultures.
24(2), 54–60. Biotechnology and Bioengineering, 110(5), 1342–1353.
10970290, 2018, 12, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/bit.26812 by CAPES, Wiley Online Library on [15/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SOMASUNDARAM ET AL. | 2907

Rameez, S., Mostafa, S. S., Miller, C., & Shukla, A. A. (2014). High‐ Tomic, S., Besnard, L., Fürst, B., Reithmeier, R., Wichmann, R., Schelling, P.,
throughput miniaturized bioreactors for cell culture process devel- & Hakemeyer, C. (2015). Complete clarification solution for proces-
opment: Reproducibility, scalability, and control. Biotechnology Pro- sing high density cell culture harvests. Separation and Purification
gress, 30(3), 718–727. https://doi.org/10.1002/btpr.1874 Technology, 141, 269–275.
Rathore, A., Kennedy, R., Donnell, J., Bemberis, I., & Kaltenbrunner, O. Udpa, N., & Million, R. P. (2016). Monoclonal antibody biosimilars. Nature
(2003). Qualification of a chromatographic column: Why and how to Reviews. Drug Discovery, 15(1), 13–14. https://doi.org/10.1038/nrd.
do it. Biopharm International, 16(3), 30–40. 2015.12
Rathore, A. S., & Sofer, G. (2012). Process validation in manufacturing of Walther, J., Godawat, R., Hwang, C., Abe, Y., Sinclair, A., & Konstantinov, K.
biopharmaceuticals (Third ed). Boca Raton, Florida: CRC Press. (2015). The business impact of an integrated continuous biomanufactur-
Rucker‐Pezzini, J., Arnold, L., Hill‐Byrne, K., Sharp, T., Avazhanskiy, M., & ing platform for recombinant protein production. Journal of Biotechnology,
Forespring, C. (2018). Single pass diafiltration integrated into a fully 213, 3–12.
continuous mAb purification process. Biotechnology and Bioengineering. Wang, A., Lewus, R., & Rathore, A. S. (2006). Comparison of different
Schaber, S. D., Gerogiorgis, D. I., Ramachandran, R., Evans, J. M. B., Barton, options for harvest of a therapeutic protein product from high cell
P. I., & Trout, B. L. (2011). Economic analysis of integrated continuous density yeast fermentation broth. Biotechnology and Bioengineering,
and batch pharmaceutical manufacturing: A case study. Industrial & 94(1), 91–104. https://doi.org/10.1002/bit.20816
Engineering Chemistry Research, 50(17), 10083–10092. https://doi.org/ Warikoo, V., Godawat, R., Brower, K., Jain, S., Cummings, D., Simons, E., …
10.1021/ie2006752 Konstantinov, K. (2012). Integrated continuous production of recom-
Scharl, T., Jungreuthmayer, C., Dürauer, A., Schweiger, S., Schröder, T., & binant therapeutic proteins. Biotechnology and Bioengineering, 109(12),
Jungbauer, A. (2016). Trend analysis of performance parameters of 3018–3029. https://doi.org/10.1002/bit.24584
pre‐packed columns for protein chromatography over a time span of Weitzhandler, M., Farnan, D., Horvath, J., Rohrer, J. S., Slingsby, R. W.,
ten years. Journal of Chromatography A, 1465, 63–70. https://doi.org/ Avdalovic, N., & Pohl, C. (1998). Protein variant separations by cation‐
10.1016/j.chroma.2016.07.054 exchange chromatography on tentacle‐type polymeric stationary
Shamashkin, M., Godavarti, R., Iskra, T., & Coffman, J. (2013). A tandem phases. Journal of Chromatography A, 828(1), 365–372. https://doi.
laboratory scale protein purification process using Protein A affinity org/10.1016/S0021‐9673(98)00521‐4
and anion exchange chromatography operated in a weak partitioning Whitford, W. G. (2014). Single‐ use systems support continuous biopro-
mode. Biotechnology and Bioengineering, 110(10), 2655–2663. cessing by perfusion culture, Continuous Processing in Pharmaceutical
Shukla, A. A., Hubbard, B., Tressel, T., Guhan, S., & Low, D. (2007). Manufacturing (183–226). Weinheim, Germany: Wiley‐VCH Verlag
Downstream processing of monoclonal antibodies—application of GmbH & Co. KGaA.
platform approaches. Journal of Chromatography B, 848(1), 28–39. Yang, W. C., Lu, J., Kwiatkowski, C., Yuan, H., Kshirsagar, R., Ryll, T., &
Shukla, A. A., Wolfe, L. S., Mostafa, S. S., & Norman, C. (2017). Evolving Huang, Y. M. (2014). Perfusion seed cultures improve biopharmaceu-
trends in mAb production processes. Bioengineering & Translational tical fed‐batch production capacity and product quality. Biotechnology
Medicine, 2(1), 58–69. https://doi.org/10.1002/btm2.10061 Progress, 30(3), 616–625.
Siu, S. C., Chia, C., Mok, Y., & Pattnaik, P. (2014). Packing of large‐scale Yang, Y., Lamm, L. J., He, P., & Kondo, T. (2002). Temperature effect on
chromatography columns with irregularly shaped glass based resins peak width and column efficiency in subcritical water chromatogra-
using a stop‐flow method. Biotechnology Progress, 30(6), 1319–1325. phy. Journal of Chromatographic Science, 40(2), 107–112.
https://doi.org/10.1002/btpr.1962 Yavorsky, D., Blanck, R., Lambalot, C., & Brunkow, R. (2003). The
Soriano, G. A., Titchener‐Hooker, N. J., & Shamlou, P. A. (1997). The clarification of bioreactor cell cultures for biopharmaceuticals (27).
effects of processing scale on the pressure drop of compressible gel Zydney, A. L. (2015). Perspectives on integrated continuous bioproces-
supports in liquid chromatographic columns. Bioprocess Engineering, sing—opportunities and challenges. Current Opinion in Chemical
17(2), 115–119. https://doi.org/10.1007/s004490050363 Engineering, 10, 8–13.
Spangler, G. E., & Collins, C. I. (1975). Peak shape analysis and plate theory Zydney, A. L. (2016). Continuous downstream processing for high value
for plasma chromatography. Analytical Chemistry, 47(3), 403–407. biological products: A Review. Biotechnology and Bioengineering, 113(3),
Steele, A., & Arias, J. (2014). Accounting for the Donnan effect in 465–475.
diafiltration optimization for high‐concentration UFDF applications.
BioProcess International, 12(1).
Steinebach, F., Ulmer, N., Wolf, M., Decker, L., Schneider, V., Wälchli, R., …
Morbidelli, M. (2017). Design and operation of a continuous SU PP ORT IN G IN FOR M ATI O N
integrated monoclonal antibody production process. Biotechnology
Progress, 33, 1303–1313. Additional supporting information may be found online in the
Stickel, J. J., & Fotopoulos, A. (2001). Pressure‐flow relationships for Supporting Information section at the end of the article.
packed beds of compressible chromatography media at laboratory
and production scale. Biotechnology Progress, 17(4), 744–751. https://
doi.org/10.1021/bp010060o
Stock, L. R., Bisschops, M., & Ransohoff, T. C. (2014). The potential impact How to cite this article: Somasundaram B, Pleitt K, Shave E,
of continuous processing on the practice and economics of biophar- Baker K, Lua LHL. Progression of continuous downstream
maceutical manufacturing, Continuous processing in pharmaceutical processing of monoclonal antibodies: Current trends and
manufacturing (479–494). Weinheim, Germany: Wiley‐VCH Verlag
challenges. Biotechnology and Bioengineering. 2018;115:
GmbH & Co. KGaA.
2893–2907. https://doi.org/10.1002/bit.26812
Thillaivinayagalingam, P., Reidy, K., Lindeberg, A., & Newcombe, A. R. (2012).
Revisiting protein A chromatography. BioProcess International, 10(3).

You might also like