Next Article in Journal
CTS-Net: A Segmentation Network for Glaucoma Optical Coherence Tomography Retinal Layer Images
Next Article in Special Issue
Downscaling Industrial-Scale Syngas Fermentation to Simulate Frequent and Irregular Dissolved Gas Concentration Shocks
Previous Article in Journal
WiFi-Based Detection of Human Subtle Motion for Health Applications
Previous Article in Special Issue
Bioprocess Economic Modeling: Decision Support Tools for the Development of Stem Cell Therapy Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

rAAV Manufacturing: The Challenges of Soft Sensing during Upstream Processing

by
Cristovão Freitas Iglesias, Jr.
1,
Milica Ristovski
1,2,
Miodrag Bolic
1 and
Miroslava Cuperlovic-Culf
3,4,*
1
Faculty of Engineering, University of Ottawa, Ottawa, ON K1N 6N5, Canada
2
Faculty of Medicine, University of Ottawa, Ottawa, ON K1H 8M5, Canada
3
Digital Technologies Research Center, National Research Council, Ottawa, ON K1A 0R6, Canada
4
Department of Biochemistry, Microbiology, and Immunology, Faculty of Medicine, University of Ottawa, Ottawa, ON K1H 8M5, Canada
*
Author to whom correspondence should be addressed.
Bioengineering 2023, 10(2), 229; https://doi.org/10.3390/bioengineering10020229
Submission received: 10 November 2022 / Revised: 31 January 2023 / Accepted: 2 February 2023 / Published: 8 February 2023
(This article belongs to the Special Issue Design, Optimization and Scale-Up of Industrial Bioprocess)

Abstract

:
Recombinant adeno-associated virus (rAAV) is the most effective viral vector technology for directly translating the genomic revolution into medicinal therapies. However, the manufacturing of rAAV viral vectors remains challenging in the upstream processing with low rAAV yield in large-scale production and high cost, limiting the generalization of rAAV-based treatments. This situation can be improved by real-time monitoring of critical process parameters (CPP) that affect critical quality attributes (CQA). To achieve this aim, soft sensing combined with predictive modeling is an important strategy that can be used for optimizing the upstream process of rAAV production by monitoring critical process variables in real time. However, the development of soft sensors for rAAV production as a fast and low-cost monitoring approach is not an easy task. This review article describes four challenges and critically discusses the possible solutions that can enable the application of soft sensors for rAAV production monitoring. The challenges from a data scientist’s perspective are (i) a predictor variable (soft-sensor inputs) set without AAV viral titer, (ii) multi-step forecasting, (iii) multiple process phases, and (iv) soft-sensor development composed of the mechanistic model.

1. Introduction

Recombinant adeno-associated viruses (rAAV) are a class of viral vectors engineered to express specific genes or therapeutic proteins in cells [1,2,3]. They are used in gene therapy as vectors to deliver genetic material to specific cells with high efficiency and low toxicity [1,4,5]. rAAVs are attractive vectors for gene therapy due to their ability to integrate their genetic load into a specific location in the host genome, remain a silent passenger in the host genome, and not cause disease [5,6,7,8]. Currently, rAAVs are being evaluated in preclinical and clinical trials for a wide range of genetic diseases [3,6,9,10], including inherited retinal diseases [11,12], muscular dystrophies [13,14,15,16], and hemophilia [17,18]. rAAVs have also shown promise in gene therapies for cancer [19], cardiovascular diseases [20,21], and neurological disorders [22,23]. The most common method for rAAV production is through transfection of host cells, such as HEK293 cells, with plasmids encoding for rAAV Rep and Cap proteins, and the therapeutic gene of interest [18,24,25]. However, for rAAVs to be used effectively in gene therapy, large amounts of high-quality virus particles are required [14,26,27]. Therefore, the production of high-quality virus particles is critical for the advancement of rAAV-based therapies.
The current state of rAAV research is focused on developing new methods to produce high-quality virus particles in large amounts [2,28,29,30,31,32,33,34,35]. Researchers have been developing new methods such as serum-free suspension systems [10,36], stable cell lines [37,38], helper-dependent AAV systems [39], high-throughput systems [40], and other methods such as using baculovirus-insect cells systems [41]. Furthermore, Srivastava et al. [42] describe the numerous manufacturing processes and difficulties faced during the manufacture and storage of rAAV and provide a road map for enhancing manufacturing workflow efficiency and extending product shelf-life. Dobrowsky et al. [43] present an overview of the manufacturing-related obstacles resulting from using this vector for clinical purposes, mainly focusing on the difficulties associated with the upstream process. However, these studies do not comment on soft sensing as a mitigation strategy to improve the scalability, reproducibility, and cost-effectiveness of rAAV production, which is crucial for commercializing rAAV-based therapies. For example, currently, a typical production run of an rAAV-vector treatment utilizing high-yield cell lines and large-capacity bioreactors may only produce ten doses of a systemic gene therapy from a single batch at the cost of about USD 100,000 per dosage (considering around 1 × 10 17 vg per batch) [44]. Consequently, any technological advancement that decreases the cost per dose would be immediately advantageous [44]. Soft sensors have been used as a fast and low-cost approach to control and monitor the bioprocesses by enabling real-time screening and control of critical process variables of the upstream process to ensure product quality and process consistency [45,46,47,48,49,50]. Furthermore, this approach is a recommendation of the Food and Drug Administration’s process analytical technology initiative (PAT). They suggest the continuous in-process monitoring of critical process variables to assure product quality at every stage of the manufacturing process [51].
The literature presentation of the usage of soft sensors for rAAV production is limited. Recently, two review articles described approaches of relevance in soft sensors in rAAV production. Tsopanoglou et al. [52] discuss the advantages and challenges of hybrid modeling for upstream pharmaceutical bioprocesses, and Joiner et al. [53] proposed an outline (high-level framework) for the future development of a hybrid mechanistic and data-driven model for rAAV production modeling. However, the main challenges of soft-sensing in rAAV production, such as the design and applications were, best to our knowledge, not considered.
Therefore, to encourage and provide a background necessary for the advancement of the field, this review article discusses the advantages and challenges of applying soft sensors as a fast and low-cost methodology for rAAV production monitoring based on current knowledge of rAAV biology and manufacturing. Furthermore, we present the possible solutions to side-step these challenges and a glossary (Table S1 in the Supplementary Material) provides common vocabulary bridging between data scientists and biologists. The challenges from a data scientist’s perspective are: (i) a predictor variable (soft sensor inputs) set without rAAV viral titer, (ii) multi-step forecasting, (iii) multiple process phases, and (iv) soft sensor development composed of the mechanistic model.
In this review, we discuss the advantages of soft-sensing in the upstream process of rAAV production in Section 2. We further discuss the soft sensor design in Section 3. The following Section 4 is concerned with the four challenges and possible solutions in soft sensor development for rAAV production from a data scientist’s perspective. Section 5 concludes with a final discussion and a summary of the topic.

2. Soft Sensor in Upstream of rAAV Production

Viral vector manufacturing is a complex process that involves an upstream and downstream component and takes into account the maintenance of the viral vector’s long-term stability and efficacy. The upstream process is comprised of four steps: plasmid development [54,55], cell expansion [56], plasmid transfection [8,57], and viral vector production [25,42]. The challenges of the upstream process of rAAV production are the need to increase the viral titer and decrease the variability in product quality [42]. These challenges in rAAV manufacturing resemble those that monoclonal antibody manufacturers faced [42]. From a biological viewpoint, two approaches can be used: (i) developing stable rAAV producer suspension cell lines that do not require triple transfection or (ii) utilizing other innovative technologies that are plasmid-free [42]. There is active development in both areas, but these are long-term solutions requiring more investigation and investment. On the other hand, mathematical modeling is a possible immediate solution from the biomanufacturing monitoring and control viewpoint since it minimizes the production time and cost while also learning about the critical process steps [52,58,59]. Mathematical models are increasingly being used in the biomanufacturing industry to design and optimize different processes and unit operations [59,60]. Mathematical modeling that can optimize the upstream process is based on the concept of soft sensors. Soft sensors use physical or machine learning models that predict or estimate quantities of interest based on raw data [61,62]. They consist of mathematical models that use data from a measurement device (input) to predict a quantity of interest (output), which is used to generate new information about the bioprocess [50,63].
Biopharma 4.0 is the next step in the digitalization of biopharmaceutical manufacturing that aims to optimize biopharmaceutical production efficiency by minimizing human intervention and the uncertainty that arises from it [52]. Soft sensors are a crucial enabler of Biopharma 4.0 since it enables real-time monitoring of variables that are difficult to measure directly or that can only be measured at a low sampling frequency [52,64]. This is a vital step towards the digital transformation of bioprocesses [52]. The application of soft sensors in upstream rAAV production can contribute to the reduction of time and production costs, the generation of relevant data, improved control of bioprocesses, increased product quality, and the reduction of raw material waste [50]. The large potential of mathematical models used in soft sensors lies in the possibility of establishing quantitative links between critical process parameters (process variables that have an impact on a critical quality attribute) and Key Performance Indicator (KPI) providing data for bioprocess optimization [50,52,65,66]. Furthermore, soft sensors are used as a base of advanced control strategies for bioprocess optimization, such as digital twins [59,67,68], model predictive control [69,70], and model-based control [45].
Soft sensors have been used to control and monitor the production of mAb in Chinese hamster ovary (CHO) cell cultures, and enable the real-time screening and control of critical process variables of the upstream process to ensure product quality and process consistency [71,72,73,74,74]. Besides applying soft sensors in mAb production, they were also developed for a number of other bioprocess applications. For example, Zhang et al. [75] used a soft sensor based on artificial neural networks to monitor the fermentation process by measuring glycerol, 1,2 propanediol, and biomass. Chromatography monitoring was performed with a soft sensor based on partial least squares regression with the help of mid-UV absorption spectra [76] as well as Fourier transform mid-infrared spectroscopy [77]. Several papers used soft sensors to estimate the biomass and substrate concentrations [78,79,80,81,82]. Krishna et al. [83] calculated the lactose and ethanol concentrations with deep learning-based soft sensor models.
Since soft sensors have helped improve quality and production rates for mAbs [71,72,73,74,74], one can expect that a similar methodology would be beneficial for rAAV production. rAAV production utilizes a mammalian cell culture-based process similar to mAb production and faces a similar upstream challenge as was faced by mAb production [42,52,84]. Therefore, soft sensors have a high potential to provide insight into the underlying micro and macro-scale phenomena of upstream bioprocesses of rAAV manufacturing and improve the yield and quality of production [52].

3. Designing Soft Sensors

Soft sensor design should combine analytical and computational methods with adequate mathematical models for efficient modeling and accurate predictions of bioprocesses. Three kinds of modeling approaches are commonly used: the Data-Driven Model (DDM), the Mechanistic Model (MM), and the Hybrid Model (HM), which aims to combine the advantages of DDMs and MMs [50].

3.1. Data Driven Models

DDMs are built purely on empirical observations of a process [50]. They establish links between input and output variables (system state variables) without having explicit knowledge of the physical laws in the system [50,85]. DDMs are mainly based on machine learning techniques that may be used to create models that can either supplement or completely replace physics-based models. DDMs learn the link between the inputs and outputs of a system from a training data set, and after the DDM has been trained, its ability to generalize to new data may be assessed using an independent data set [50,85]. Some of the most common data-driven modeling techniques used in related applications are: fuzzy rule-based systems [85], Partial Least Squares (PLS) [76,77], Artificial Neural Networks [75], Support Vector Machine (SVM) [85], and Deep Learning (DL) [86]. In the last two decades, DL models composed of multiple hidden layers have been shown to be capable of learning data representations with various levels of abstraction [86]. As a result, these methods have dramatically improved many domains, such as drug discovery and genomics [87]. Research that uses DL methods in soft sensors has been gaining popularity in recent years [86]. The advantages of DL approaches over conventional algorithms mainly result from (i) learning relevant process behaviors without knowledge or experience about process mechanisms and (ii) taking full advantage of the massive amount of data available in public data sets for performance improvements [49,86,88]. Typical models, i.e., architecture, in the DL field for soft sensors include Autoendocoder (AE), Convolutional Neural Networks (CNN), and Recurrent Neural Networks (RNN), such as Long Short Term Memory (LSTM) networks [86].
DDM Advantages. DDMs, compared to kinetic models, have more parameters and structures for data regression, allowing the integration of various process behaviors obtained under different operating circumstances in a single model. This is done using a training set that can represent all possible behaviors seen in the system [89]. In addition, DDMs can identify patterns and trends in complex and heterogeneous data, thus facilitating informed decisions and actions without insight into the fundamental mechanism of the process [52,90]. Another important advantage of DDMs is that the development time is shorter than that of MM design [52].
DDM Disadvantages. Unfortunately, DDMs have several significant disadvantages. For example, complex DDMs, such as deep learning models, perform predictions without consideration of the underlying mechanisms, resulting in parameters with no physical, chemical, or biological meaning. The variables are based on correlations and not necessarily on causation [52]. Furthermore, poor data quality and low data quantity are a major concern for DDMs. The variables can only be predicted with high accuracy in the space adequately represented by the training set. This means that generalization ability relies on training with large amounts of data [52,90]. Although designing DDMs is usually less time-consuming than designing an MM [52,91], DDMs may require pre-processing steps, which can be time-consuming [52]. DDMs need accessible, representative, and reliable training data sets. DDMs can have limited bioprocess control and optimization capabilities and frequent model drift (the degradation of the prediction power of the model due to the changes in the statistical properties of the input features, target variable, or relationships among variables). Re-calibration is often time-consuming [50] and data demanding.
DDM in upstream rAAV production. DDMs are often applied in academia and industry, particularly in the chemical process sectors [50]. An initial study using DDMs for the upstream processing of rAAVs was recently published. Iglesias et al. [92] developed an approach for handling a massive proportion of missing labels in multivariate, multi-step time series forecasting. The approach is a two-step process where interpolation (using Gaussian Processes Regression (GPR) and domain knowledge from experts) and prediction modeling are separated to integrate the prior domain knowledge. The authors evaluated their approach by developing a conceptual soft sensor to forecast the biomass of HEK293 cells in rAAV production. The soft sensor was designed using Long Short-Term Memory encoder-decoder architecture to forecast the biomass in the next 1 h based on measuring the last hour of three input features representing the model’s input. The input features include: (i) Cumulative O2 (mL), (ii) Dissolved Oxygen (%), and (iii) Cumulative Dissolved Oxygen. The model was trained and tested with a real data set obtained during rAAV production performed in 3L bioreactors [92].

3.2. Mechanistic Models

Physical, chemical, and biological insights and understanding of the transformation processes in producing organisms are used in MMs development. The recommended mechanistic technique for capturing the underlying events during cell culture is kinetic modeling [50,52,93]. Mechanistic, white-box or parametric models are considered a priori, based on the knowledge of the process. The number of model parameters is fixed, and they might have a physical or empirical interpretation depending on the level of knowledge.
Usually, MMs are represented as systems of ordinary or partial differential equations, and the numerical values are assigned to the MMs parameters via parameters estimation [94,95,96,97,98] (see Section S1 in the Supplementary Materials). Two approaches that can be used are Neural Ordinary Differential Equation (NODE) [99,100,101,102,103,104,105,106] and Bayesian Inference [107,108] (see Section S1 in the Supplementary Materials). Furthermore, MMs can be categorized as structured and unstructured models [52,93].
Unstructured models treat all cells as black boxes, assuming they are all in the same physiological condition. Extracellular factors are the only ones that impact cell growth and behavior. Unstructured models rely on nutrient and metabolite concentrations in cultures while ignoring intracellular activities. In mammalian cell cultures, unstructured kinetic models have also been utilized to estimate recombinant protein output and can be used in the upstream rAAV production [93,109]. The Monod model, which assumes that the concentration of a substrate is the growth limiting factor, is a commonly used unstructured model [93]. For example, Kornecki et al. [84] used a Monod-type process model for the simulation of dynamic cellular states (i.e., lag, exponential, stationary, decline phase), as well as the uptake of substrates (i.e., glucose, glutamine), production of metabolites (i.e., lactate, ammonium), and the product (i.e., monoclonal antibody). Because of the dynamic behavior of the culture, yield coefficients can be quantitatively defined by systematically identifying cellular phases (i.e., lag, exponential, stationary, decrease). In addition, cell-dependent half-maximum rates influence the growth rate, and the model parameters were based on maximal growth rates at various substrate concentrations. They were not susceptible to change during culture.
A structured model expresses intracellular activities in structural and physiological terms, providing the most accurate picture of a cell [93,109]. Biomass is divided into sections with diverse functions in structured models. As a result, cells are not considered homogeneous, increasing the complexity of the models and the number of model parameters. It is essential to point out that, although changes in cell attributes may be described, it is very commonly assumed that all cells are the same and change at the same rate. In addition, most kinetic models in the literature are unstructured, and a structured model is not appropriate for most cell lines [93,109].
Structured and unstructured models are further divided into segregated and non-segregated models. Segregated models take into account cell populations when not all cells are homogeneous, non-segregated models consider that all cells are in an identical state [52,93]. Segregation refers to cell population heterogeneity. Similarly to structure, if the model is non-segregated, it does not have this information, and if it is segregated, it has the information [52]. Segregated models are more complex than non-segregated models. Their complexity makes parameter identification and the utilization of optimization algorithms challenging and time-consuming [93].
MM Advantages. The advantage of MMs is that they aim to represent process knowledge and generally show a higher extrapolation power than DDMs, leading to more applications. Once they are developed, MMs are reliable and reproducible, offering accurate and relevant information about the process. MMs are based on mechanistic knowledge in biology, chemistry, and physics and yield enhanced process understanding. MMs are an excellent candidate for optimization and can be used for model-based design of optimally informative experiments.
MM Disadvantages. The development of MMs may take longer than that of DDMs. The MMs require expertise in the application domain. If an MM is too simple, it may have limited extrapolation power [91]. On the other hand, if there are too many parameters, then the model is hard to train and may lack robustness and universality [52]. Although MMs seem to provide the most accurate and in-depth information, the main disadvantage is that extensive experimental effort is required for model validation. Due to the requirement for in-depth knowledge of the process, it may not be easy to automate and formalize model assembly. Finally, MMs may be difficult to deploy within the industry because of the high level of expertise required for operation and calibration [50]. When there is a limited understanding of underlying intracellular mechanisms and regulation between process inputs and outputs in mammalian cell cultures, the calibration of an MM is a difficult task [94,95].
MM in upstream rAAV production. We are at the beginning of the mathematical modelling for rAAV production. Nguyen et al. [110] presented the first structured MM for rAAV production. The authors proposed a mechanistic model for synthesizing rAAV viral vectors by triple plasmid transfection based on the underlying biological processes gathered from wild-type rAAV to better understand the dynamics of recombinant viral generation. The model includes essential phases such as exogenous DNA delivery, the reaction cascade that produces viral proteins and DNA, resulting in filled capsids, and the Rep protein’s complex activities as a regulator of packing plasmid gene expression and a viral DNA packaging catalyst. This model has demonstrated that it can provide valuable insights into experimental data, point out process bottlenecks, and direct future studies. On the other hand, recently, Iglesias et al. [111] presented the first Extended Kalman Filter (EKF) that uses an unstructured mechanistic model as a process model to monitor rAAV production. The proposed approach uses only online viable cell density (Xv) measurements to estimate the other process state variables, including the concentrations of glucose (GLC), glutamine (GLN), lactate (LAC), ammonium (AMM), and rAAV viral titers measured at a low sampling frequency.

3.3. Hybrid Models

Hybrid or grey-box models combine nonparametric and parametric models that are based on different types of knowledge [112,113,114]. Furthermore, the vast majority of hybrid modeling studies explored the combination of MM and DDM [115]. The information available through mechanistic modeling is the a priori knowledge in the form of material and energy balances, thermodynamics, and kinetic laws. In addition, mechanistic relationships, such as physical, chemical, or biochemical kinetics, are covered by mechanistic models [93]. Data-driven knowledge, in the form of heuristics, is the second type of knowledge. For example, heuristic process knowledge (e.g., thresholds for inhibition, limitation or optimal temperature, pH, DO ranges) is frequently expressed in fuzzy sets, and historical data are transformed into models by neural networks [93].
Hybrid modeling compromises the benefits and drawbacks of MMs and DDMs. Tsopanoglou et al. [52] state that HMs may lead to safer, more effective, personal, and low-cost products. While MMs can provide predicted outcomes, their construction is time-consuming and needs comprehensive process understanding. DDMs, on the other hand, are easy to construct and use, but they only have acceptable descriptive features within the range of the data they are based on. Compared to solely data-driven modeling techniques, the major benefit of hybrid models is that they can achieve higher accuracy, more efficient model creation, and better extrapolation features [93]. The first challenge in developing HM is figuring out the right way to combine MM and DDM for a specific application, and this can be done in parallel or sequential (cascade) ways [112,116]. In addition, one disadvantage of HM is that the implementation and their parameterization algorithms are highly prone to errors and are time-consuming [112,116].
HM in upstream rAAV production. Hybrid models have been successfully applied in bioprocess control, monitoring, and optimization in bioreactors. Therefore, they could be utilized to design soft sensors [50,93] but, so far, no HM has been developed and reported to rAAV production monitoring in the upstream process.

4. Challenge of Soft Sensors in Upstream rAAV Production and Possible Solutions

Since soft sensors are sometimes the only way to determine critical process parameters (CPP) or critical quality attributes (CQA) online, they have grown in importance within quality by design (QbD) and process analytical technology initiatives. However, data scientists face several significant challenges in developing soft sensors for rAAV production: (i) Predictor variable (soft sensor inputs) set without rAAV viral titer, (ii) Multi-step forecasting, (iii) Multiple process phases, and (iv) soft sensor development composed of MMs.

4.1. Predictor Variables (Soft Sensor Inputs) Set without rAAV Viral Titer

The current techniques for monitoring rAAV manufacturing in bioreactors are expensive, laborious, and time-consuming. The measure of the CPPs, such as cell density and metabolites, and the quantification of the CQAs, such as rAAV genome titer by quantitative polymer chain reaction (qPCR)/Droplet Digital polymerase chain reaction (ddPCR) or viral capsid titer by Enzyme-linked immunosorbent assay (ELISA), takes one day to complete and requires sampling [117]. Recently, in situ monitoring technologies, such as Raman spectroscopy [118,119] and fluorescence spectroscopy [120,121,122], have been developed to estimate the cell density and measure metabolite concentrations in mammalian cell cultures in real-time but there were no reports thus-far on using them to estimate rAAV titer. Additionally, the setup of a spectroscopic system is expensive in terms of both financial outlay and calibration work [123]. Therefore, in most cases, rAAV viral titer measurements are unavailable during production. So, the challenge is to design a soft sensor that does not rely on rAAV viral titer measurements as its input. A soft sensor can have cell density or/and metabolite concentration (i.e., GLC, GLN, LAC, and AMM) as input and the rAAV viral titer only as output. For example, it is impossible to use past rAAV viral titer measurements to predict the present/future state because they are unavailable during the rAAV manufacturing process. In most cases, the quantification of viral titer in rAAV production is done only at the end of production. After the production is completed, the samples collected are used to quantify the viral titer, and the ddPCR takes 1 day to complete the process. Therefore, during the rAAV manufacturing, we have only cell density, GLC, GLN, LAC, and AMM that are measured offline every ∼30 min, or cell density that can be easily measured online.
Possible Solution. A soft sensor developed with a DDM, such as an LSTM, can learn how to predict the rAAV viral titer (CQA) based only on CPP, such as cell density, GLC, GLN, LAC, and AMM. More details about the development of an LSTM can be found in [92,124].

4.2. Multi-Step Forecasting

Time-series forecasting models predict future values of one or more dependent variables based on the past values of one or more independent variables [124,125]. Time series models can be univariate (with only one time-dependent or independent variable) or multivariate (with several time-dependent and independent variables). Although there are significant differences between univariate and multivariate models, most deep-learning models can handle both. For example, the model can have multivariate time series as inputs and univariate time series as the output [126,127]. Time series models can be defined with regard to the forecasting horizon in the n-step-ahead term. For example, if the model forecasts a point that is a one-time step ahead in the future, it is denoted as a one-step-ahead forecast. Similarly, multi-step forecast predicts n times-steps ahead [126,127]. In the majority of soft sensors applied in bioprocess applications, the one-step multivariate forecast is the most common approach used to estimate biomass and viral titer [49,86,128,129,130,131,132,133,134,135,136,137]. It is standard practice when forecasting the biomass to use lagged observations t 1 time instant as input variables to forecast the variables at the current time step t. An effective strategy for improving the process monitoring capability is to predict rAAV viral titer during the upstream process to assess future bioprocess conditions. It would be helpful to forecast the time of transfection and harvest (the time to stop the bioreactor after the production achieves the plateau). However, this would require a multivariate multi-step forecasting soft sensor, and the challenge here is to design a soft sensor to multi-step forecast the viral titer 24 or 48 h ahead. Unfortunately, this is difficult because of the accumulation of errors, reduced accuracy, and increased uncertainty.
Possible Solution. A soft sensor developed with a DDM, such as an LSTM, can be trained/tested to multi-step forecast a CQA based on previous time series data of CPPs [92]. LSTM is capable of capturing the patterns of long term, and this can provide greater accuracy. Furthermore, an ensemble approach could be used in LSTM development to perform forecasts with uncertainty [138]. Figure 1 shows a possible approach for data-driven soft sensor development for multi-step forecasting of rAAV viral titer. This approach aims to deploy unstructured MMs to generate training data sets for a DDM in order to learn to forecast (multi-step) CPP and CQA in rAAV production, including viable cell density (Xv) and rAAV viral titer, the two main critical process variables, see Figure 1. They are essential because the timing of transfection and harvest procedures, as well as monitoring the overall status of the culture, depends on the cell concentration. Furthermore, forecasting the rAAVviral titer is particularly interesting to the pharmaceutical manufacturing industry because it is a CQA. The first stage of this approach starts with the design of experiments (DoE) [116,139] to systematically investigate the relationship between CPPs (input factors such as Xv, GLC, GLN, LAC, and AMM) and rAAV viral titer (CQA) to have a better understanding of the impacts of different levels of each CPPs (different cell culture conditions) on the rAAV viral titer. In this stage, the measurements can be online for Xv, GLC, GLU, LAC, and AMM using spectroscopy-based techniques [64,118,119,120,121,122,140,141,142,143,144,145,146], and offline for rAAV viral titer using qPCR/ddPCR or ELISA, and they are stored in a data storage. These measurements from the DoE represent different cell culture conditions for rAAV production and are obtained with missing data and at different scales, since, for example, offline data are not collected at the same frequency as online data. Therefore, they should be integrated to generate an information-rich dataset (to enable effective DDM training), possibly performed with an unstructured MM. The measurements of data storage can be used for the parameter estimation of an unstructured MM regards the different rAAV production conditions, and it can generate a training dataset (information-rich) to enable DDMs such as LSTM models to learn to forecast the Xv and rAAV viral titer on different cell culture conditions of rAAV production. After the training, this DDM can be used in the production stage to perform a multi-step forecast of the timing of transfection and harvest procedures based on the CPP, and CQA estimation obtained from online measurements, see Figure 1. More details about the development of an LSTM can be found in [92,124].
This approach is based on [52,92,147]. It would not improve process comprehension regarding rAAV production, but it could help apply soft sensors in bioreactors for process optimization by efficiently forecasting a CQA, such as rAAV viral titer.

4.3. Multi-Phase Process

The upstream process of rAAV production is a multi-phase process composed of cell expansion and viral vector production. In the cell expansion phase, the cells are multiplied until they reach the required cell density to start the plasmid transfection, and in the viral vector production, the virus is produced for many days in the transiently transfected cells. Therefore, in developing a soft sensor methodology, it is essential to consider multi-phase characteristics of the upstream process in rAAV production since the relationships between state process variables (which can be physical, chemical, and biological) can vary substantially in the individual phases. Online monitoring should furthermore generate information regarding transitory process states [50]. Then, monitoring these phases can generate essential information that allows the tracking and fine-tuning of crucial process parameters in the upstream process, resulting in more efficient control, optimal harvest timing, and enhanced end-product quality.
Possible Solution. There are two potential directions. The first strategy would be to divide the process data sets into individual phase segments and develop separate sub-models for these segments [63]. In this approach, the overall model comprises one multi-variable sub-model for the cell expansion and the other for the viral vector production phase. A similar approach was developed by Selişteanu et al. [148] to model Monoclonal Antibody Production, where they created sub-models corresponding to the dynamics of viable cell concentration and dead cell concentration that are used in the time period of the exponential growth phase. A clear definition of the phases is a challenge in a multiple-model framework since biological effects can overlap [93]. However, algorithms for phase detection can be based on the shape of process trajectories or the correlation structure of process variables. For the detection and division of process phases, trajectory-based and correlation-based methods have been proposed in the literature [63]. The second approach involves modeling the entire upstream process rather than just one phase, considering interactions between phases or/and their relative relevance to the final product quality [149]. This method, known as multi-block, divides vast data collections by the process state variables into functional blocks, typically one block per phase. Then, the blocks are combined to most accurately forecast product state variables. An example of multi-block modeling is multi-block partial least squares [149].

4.4. Soft Sensor Development Composed of MMs

The soft sensor design based on HM, composed of MM and DMM, could be a challenging task in the case of rAAV production in bioreactors. The main concern is with MM, since the current state-of-the-art knowledge related to viral vector manufacturing modeling is still limited. There are relatively few works that focus on critical factors in rAAV manufacturing processes [53,110,111]. The first MM for rAAV production was published only recently by Nguyen et al. [110], and it is a structured MM. The model describes the kinetic behavior of transient transfection at the sub-cellular level and does not enable the macro-modeling of rAAV production at a bioreactor level (macro-scale) [116]. Furthermore, the 21 variables (species) available in this structured mechanistic model are not commonly measured in the bioreactor as bioprocess parameters, such as cell density, glucose (GLC), lactate (LAC), and oxygen [64]. This reduces the applicability of this structured MM to process optimization and the control of rAAV production at the bioreactor level. To our knowledge, there is currently no HM soft sensor developed and reported for rAAV production monitoring besides a proposal for HM development published recently by Joiner et al. [53]. The authors proposed a framework at a very abstract level that can be used as the initial direction for developing a hybrid mechanistic and data-driven model for rAAV production. However, this proposal suggests incorporating all mechanistic aspects of HEK293 metabolism and rAAV into a single HM, and this could be a very challenging development due to process complexity when involving structured and unstructured mechanistic models (multiscale modeling) [116] and a lack of information about the kinetic parameters in the literature.
Possible solution. One solution to facilitate the hybrid soft sensor development is to adapt unstructured MM from the literature since, most often, operational and control processes for optimal bioreactor performance can be derived from unstructured MM [150]. For example, unstructured MMs have been widely employed in monoclonal antibody production using Chinese hamster ovary (CHO) cells, and this approach could be used in rAAV production monitoring to speed up the optimization of the upstream process as well [2]. In this sense, the unstructured MM proposed by Kornecki et al. [84], which is also based on mammalian cell cultures in rAAV production, is a good starting point from which to design a hybrid soft sensor for rAAV production monitoring without considering the complexity of a triple plasmid transfection process. Therefore, this approach can help apply soft sensors in bioreactors for rAAV production monitoring, even though it cannot improve the understanding of the rAAV production mechanism.
Figure 2 shows a possible approach for developing a soft sensor based on HM, considering an unstructured MM as the sub-model. In this development approach, an HM for real-time monitoring of rAAV viral titer could be developed by combining an extended Kalman filter and an artificial neural network (ANN), see Figure 2. The EKF requires process and measurement models to estimate a process’s state variables and, generally, an unstructured MM is used as a process model. For example, the EKF proposed by Iglesias et al. [111] could reduce the number of devices for monitoring the state variables and enable the online monitoring of the rAAV viral titer. However, the parameter estimation for the process model (an unstructured MM) performed by the joint approach in EKF should be improved because it does not guarantee convergence. Alternatively, another method can estimate parameters outside the EKF calculation, such as an ANN. Therefore, the first stage of this approach would be training and testing an ANN to estimate the parameters of unstructured MM used as a process model of EKF, and this can be done similarly to the first stage of the strategy for soft sensor development based on DDM described in Section 4.2, see Figure 1. The difference is that here, the DDM (ANN) only estimates the parameters θ of the EKF process model. See Figure 2. After the training, this ANN model can be used in the production stage to work with the EKF to perform real-time estimations of rAAV viral titer using the online measurements of CPP such as Xv, GLC, LAC, and AMM; see Figure 2.
This approach is based on [111]. It would not improve process comprehension regarding rAAV production, but it might help apply soft sensors in bioreactors for process optimization by efficiently real-time monitoring a CQA, such as rAAV viral titer. It is important to point out that hybrid models involving Kalmen filter methods were proposed in [151,152] for the real-time monitoring and control of mammalian cell culture and presented significant results. More details about EKF can be found in [111].

5. Conclusions

Recombinant adeno-associated virus (rAAV) has rapidly emerged as one of the most attractive viral transfer tools, and several new advances have been made in different aspects of rAAV production. However, it is still a nascent gene therapy field that demands robust, scalable, and economically viable manufacturing processes. Therefore, optimizing upstream processing of rAAV production is extremely important to maximize viral titer and find stability for higher product quality, and soft sensors can be a valuable tool for this task. Furthermore, the cost-effectiveness of rAAV production is crucial for the commercialization of rAAV-based therapies, and soft sensing can play a key role in improving this cost-effectiveness by providing fast and low-cost monitoring of rAAV production. The others advantages of using soft sensors include productivity improvement, and product quality stabilization, which are the typical needs and goals of a manufacturing company. Despite the limited literature on the use of soft sensors in rAAV production, this review article discussed the advantages and challenges of applying soft sensors in this area and provided possible solutions to overcome these challenges. The challenges and requirements for the soft sensor development in upstream AAV production are: (i) Identifying the appropriate predictor variables (soft sensor inputs) without the use of rAAV viral titer; (ii) Design of a soft sensor for multi-step forecasting, including the time of transfection and harvest; (iii) Handling of the multi-phase characteristic of the upstream process; and (iv) Design of soft sensor composed of MMs. In the Biopharma 4.0 era, smart bio-manufacturing supported by soft sensors will produce widely accessible biopharmaceuticals that are safer, more effective, personalized, and cost-efficient. Additionally, sophisticated techniques, such as digital twins for the bioreactor process, may find a foundation in soft sensor technology.

Supplementary Materials

The following are available at https://www.mdpi.com/article/10.3390/bioengineering10020229/s1, Table S1: Glossary.

Author Contributions

Conceptualization, C.F.I.J.; writing—original draft preparation, M.R. and C.F.I.J.; writing—review and editing, M.C.-C. and M.B.; funding acquisition, M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Council through AI for Design Challenge Program (operating grants AI4D-103-1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank Create-Best for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cole, L.; Fernandes, D.; Hussain, M.T.; Kaszuba, M.; Stenson, J.; Markova, N. Characterization of recombinant adeno-associated viruses (rAAVs) for gene therapy using orthogonal techniques. Pharmaceutics 2021, 13, 586. [Google Scholar] [CrossRef] [PubMed]
  2. Sha, S.; Maloney, A.J.; Katsikis, G.; Nguyen, T.N.; Neufeld, C.; Wolfrum, J.; Barone, P.W.; Springs, S.L.; Manalis, S.R.; Sinskey, A.J.; et al. Cellular pathways of recombinant adeno-associated virus production for gene therapy. Biotechnol. Adv. 2021, 49, 107764. [Google Scholar] [CrossRef] [PubMed]
  3. Naso, M.F.; Tomkowicz, B.; Perry, W.L.; Strohl, W.R. Adeno-Associated Virus (AAV) as a Vector for Gene Therapy. BioDrugs 2017, 31, 317–334. [Google Scholar] [CrossRef] [PubMed]
  4. De Alencastro, G.; Pekrun, K.; Valdmanis, P.; Tiffany, M.; Xu, J.; Kay, M.A. Tracking Adeno-Associated Virus Capsid Evolution by High-Throughput Sequencing. Hum. Gene Ther. 2020, 31, 553–564. [Google Scholar] [CrossRef]
  5. Penaud-Budloo, M.; François, A.; Clément, N.; Ayuso, E. Pharmacology of Recombinant Adeno-associated Virus Production. Mol. Ther.-Methods Clin. Dev. 2018, 8, 166–180. [Google Scholar] [CrossRef]
  6. Clément, N.; Grieger, J.C. Manufacturing of recombinant adeno-associated viral vectors for clinical trials. Mol. Ther.-Methods Clin. Dev. 2016, 3, 16002. [Google Scholar] [CrossRef]
  7. Pillay, S.; Meyer, N.L.; Puschnik, A.S.; Davulcu, O.; Diep, J.; Ishikawa, Y.; Jae, L.T.; Wosen, J.E.; Nagamine, C.M.; Chapman, M.S.; et al. An essential receptor for adeno-associated virus infection. Nature 2016, 530, 108–112. [Google Scholar] [CrossRef]
  8. Ronzitti, G.; Gross, D.A.; Mingozzi, F. Human Immune Responses to Adeno-Associated Virus (AAV) Vectors. Front. Immunol. 2020, 11, 670. [Google Scholar] [CrossRef]
  9. Wang, D.; Tai, P.W.L.; Gao, G. Adeno-associated virus vector as a platform for gene therapy delivery. Nat. Rev. Drug Discov. 2019, 18, 358–378. [Google Scholar] [CrossRef]
  10. Adamson-Small, L.; Potter, M.; Byrne, B.J.; Clément, N. Sodium Chloride Enhances Recombinant Adeno-Associated Virus Production in a Serum-Free Suspension Manufacturing Platform Using the Herpes Simplex Virus System. Hum. Gene Ther. Methods 2017, 28, 1–14. [Google Scholar] [CrossRef] [Green Version]
  11. Giles, A.R.; Govindasamy, L.; Somanathan, S.; Wilson, J.M. Mapping an Adeno-associated Virus 9-Specific Neutralizing Epitope To Develop Next-Generation Gene Delivery Vectors. J. Virol. 2018, 92, e01011-18. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, S.K.; Lapan, S.W.; Hong, C.M.; Krause, T.B.; Cepko, C.L. In Situ Detection of Adeno-associated Viral Vector Genomes with SABER-FISH. Mol. Ther. Methods Clin. Dev. 2020, 19, 376. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, L.; Li, J.; Xiao, X. Directed evolution of adeno-associated virus (AAV) as vector for muscle gene therapy. Methods Mol. Biol. 2011, 709, 127–139. [Google Scholar] [CrossRef]
  14. Adamson-Small, L.; Potter, M.; Falk, D.J.; Cleaver, B.; Byrne, B.J.; Clément, N. A scalable method for the production of high-titer and high-quality adeno-associated type 9 vectors using the HSV platform. Mol. Ther.-Methods Clin. Dev. 2016, 3, 16031. [Google Scholar] [CrossRef] [PubMed]
  15. Bowles, D.E.; McPhee, S.W.; Li, C.; Gray, S.J.; Samulski, J.J.; Camp, A.S.; Li, J.; Wang, B.; Monahan, P.E.; Rabinowitz, J.E.; et al. Phase 1 gene therapy for Duchenne muscular dystrophy using a translational optimized AAV vector. Mol. Ther. 2012, 20, 443–455. [Google Scholar] [CrossRef]
  16. Denard, J.; Rouillon, J.; Leger, T.; Garcia, C.; Lambert, M.P.; Griffith, G.; Jenny, C.; Camadro, J.M.; Garcia, L.; Svinartchouk, F. AAV-8 and AAV-9 Vectors Cooperate with Serum Proteins Differently Than AAV-1 and AAV-6. Mol. Ther.-Methods Clin. Dev. 2018, 10, 291–302. [Google Scholar] [CrossRef]
  17. Tang, F.; Wong, H.; Ng, C.M. Rational Clinical Dose Selection of Adeno-Associated Virus-Mediated Gene Therapy Based on Allometric Principles. Clin. Pharmacol. Ther. 2021, 110, 803–807. [Google Scholar] [CrossRef]
  18. Grieger, J.C.; Soltys, S.M.; Samulski, R.J. Production of recombinant adeno-associated virus vectors using suspension HEK293 cells and continuous harvest of vector from the culture media for GMP FIX and FLT1 clinical vector. Mol. Ther. 2016, 24, 287–297. [Google Scholar] [CrossRef]
  19. Bower, J.J.; Song, L.; Bastola, P.; Hirsch, M.L. Harnessing the natural biology of Adeno-associated virus to enhance the efficacy of cancer gene therapy. Viruses 2021, 13, 1205. [Google Scholar] [CrossRef]
  20. Lu, D.; Cushman, S.; Thum, T.; Bär, C. Gene Therapy and Cardiovascular Diseases. In Genome Editing in Cardiovascular and Metabolic Diseases; Springer: Berlin/Heidelberg, Germany, 2022; pp. 235–254. [Google Scholar]
  21. Md, S.A. Cardiac disease: Current approaches to gene therapy. Bull. Med. Clin. Res. 2020, 1, 62–75. [Google Scholar]
  22. Flotte, T.R.; Gessler, D.J. Gene therapy for rare neurological disorders. Clin. Pharmacol. Ther. 2022, 111, 743–757. [Google Scholar] [CrossRef] [PubMed]
  23. Staneka, L.M.; Shihabuddina, L.S.; Chenga, S.H. Gene therapy of neurological diseases. In Advanced Textbook on Gene Transfer, Gene Therapy and Genetic Pharmacology: Principles, Delivery and Pharmacological and Biomedical Applications of Nucleotide-Based Therapies; World Scientific: Singapore, 2019; p. 471. [Google Scholar]
  24. Robert, M.A.; Chahal, P.S.; Audy, A.; Kamen, A.; Gilbert, R.; Gaillet, B. Manufacturing of recombinant adeno-associated viruses using mammalian expression platforms. Biotechnol. J. 2017, 12, 1600193. [Google Scholar] [CrossRef] [PubMed]
  25. Rumachik, N.G.; Malaker, S.A.; Poweleit, N.; Maynard, L.H.; Adams, C.M.; Leib, R.D.; Cirolia, G.; Thomas, D.; Stamnes, S.; Holt, K.; et al. Methods Matter: Standard Production Platforms for Recombinant AAV Produce Chemically and Functionally Distinct Vectors. Mol. Ther.-Methods Clin. Dev. 2020, 18, 98–118. [Google Scholar] [CrossRef] [PubMed]
  26. Clément, N. Large-scale clinical manufacturing of AAV vectors for systemic muscle gene therapy. In Muscle Gene Therapy; Springer: Berlin/Heidelberg, Germany, 2019; pp. 253–273. [Google Scholar]
  27. Ayuso, E. Manufacturing of recombinant adeno-associated viral vectors: New technologies are welcome. Mol. Ther.-Methods Clin. Dev. 2016, 3, 15049. [Google Scholar] [CrossRef]
  28. Zhao, H.; Lee, K.J.; Daris, M.; Lin, Y.; Wolfe, T.; Sheng, J.; Plewa, C.; Wang, S.; Meisen, W.H. Creation of a high-yield AAV vector production platform in suspension cells using a design-of-experiment approach. Mol. Ther.-Methods Clin. Dev. 2020, 18, 312–320. [Google Scholar] [CrossRef] [PubMed]
  29. Wosnitzka, K.; Pariag, S.; Trottin, H.; Nyamay’Antu, A.; Porte, M.; Hellal, M.; Erbacher, P. Optimization of AAV process development: Transfection matters. Cell. Gene Ther. Insights 2021, 7, 1–7. [Google Scholar] [CrossRef]
  30. Sanderson, T.; Erlandson, T.; Hazi, N.; MacIntyre, A.; Ingersoll, B.I.; McLaughlin, M.; Wesel, S.; Maples, P.B. Scalability comparison between 50 and 500 liter stirred tank bioreactor for production of rAAV viral vector. Cell Gene Ther. Insights 2021, 7, 1025–1033. [Google Scholar] [CrossRef]
  31. Zhao, H.; Meisen, H.; Wang, S.; Lee, K.J. Process Development of rAAV Production Platform Results in High Production Yield and Purity. Human Gene Therapy 2022, 34. [Google Scholar] [CrossRef]
  32. Pais, D.A.; Brown, C.; Neuman, A.; Mathur, K.; Isidro, I.A.; Alves, P.M.; Slade, P.G. Dielectric spectroscopy to improve the production of raav used in gene therapy. Processes 2020, 8, 1456. [Google Scholar] [CrossRef]
  33. Rieser, R.; Koch, J.; Faccioli, G.; Richter, K.; Menzen, T.; Biel, M.; Winter, G.; Michalakis, S. Comparison of different liquid chromatography-based purification strategies for adeno-associated virus vectors. Pharmaceutics 2021, 13, 748. [Google Scholar] [CrossRef]
  34. Guan, J.S.; Chen, K.; Si, Y.; Kim, T.; Zhou, Z.; Kim, S.; Zhou, L.; Liu, X. Process Improvement of Adeno-Associated Virus Production. Front. Chem. Eng. 2022, 4, 1. [Google Scholar] [CrossRef]
  35. Joshi, P.R.; Venereo-Sanchez, A.; Chahal, P.S.; Kamen, A.A. Advancements in molecular design and bioprocessing of recombinant adeno-associated virus gene delivery vectors using the insect-cell baculovirus expression platform. Biotechnol. J. 2021, 16, 2000021. [Google Scholar] [CrossRef]
  36. Jang, M.; Pete, E.S.; Bruheim, P. The impact of serum-free culture on HEK293 cells: From the establishment of suspension and adherent serum-free adaptation cultures to the investigation of growth and metabolic profiles. Front. Bioeng. Biotechnol. 2022, 10, 964397. [Google Scholar] [CrossRef]
  37. Lee, Z.; Lu, M.; Irfanullah, E.; Soukup, M.; Hu, W.S. Construction of an rAAV Producer Cell Line through Synthetic Biology. ACS Synth. Biol. 2022, 11, 3285–3295. [Google Scholar] [CrossRef] [PubMed]
  38. Shupe, J.; Zhang, A.; Odenwelder, D.C.; Dobrowsky, T. Gene therapy: Challenges in cell culture scale-up. Curr. Opin. Biotechnol. 2022, 75, 102721. [Google Scholar] [CrossRef] [PubMed]
  39. Weger, S. High-Level rAAV Vector Production by rAdV-Mediated Amplification of Small Amounts of Input Vector. Viruses 2022, 15, 64. [Google Scholar] [CrossRef] [PubMed]
  40. Alcudia, J.; Wright, J.F. Viral vector manufacturing: Quality attributes of rAAV used in clinical development. J. Biomol. Tech. JBT 2020, 31, S43. [Google Scholar]
  41. Wu, Y.; Han, Z.; Duan, M.; Jiang, L.; Tian, T.; Jin, D.; Wang, Q.; Xu, F. Popularizing recombinant baculovirus-derived onebac system for laboratory production of all recombinant adeno-associated virus vector serotypes. Curr. Gene Ther. 2021, 21, 167–176. [Google Scholar] [CrossRef]
  42. Srivastava, A.; Mallela, K.M.; Deorkar, N.; Brophy, G. Manufacturing challenges and rational formulation development for AAV viral vectors. J. Pharm. Sci. 2021, 110, 2609–2624. [Google Scholar] [CrossRef]
  43. Dobrowsky, T.; Gianni, D.; Pieracci, J.; Suh, J. AAV manufacturing for clinical use: Insights on current challenges from the upstream process perspective. Curr. Opin. Biomed. Eng. 2021, 20, 100353. [Google Scholar] [CrossRef]
  44. Capra, E.; Godfre, A.; Loche, A.; Smith, J. Gene-Therapy Innovation: Unlocking the Promise of Viral Vectors. 2021. Available online: https://www.mckinsey.com/industries/life-sciences/our-insights/gene-therapy-innovation-unlocking-the-promise-of-viral-vectors (accessed on 1 February 2023).
  45. Rathore, A.S.; Mishra, S.; Nikita, S.; Priyanka, P. Bioprocess control: Current progress and future perspectives. Life 2021, 11, 557. [Google Scholar] [CrossRef] [PubMed]
  46. López, R.A.; Pérez, P.A.L.; Femat, R. Control in Bioprocessing: Modeling, Estimation and the Use of Soft Sensors; John Wiley & Sons: Hoboken, NJ, USA, 2020. [Google Scholar]
  47. Vaitkus, V.; Brazauskas, K.; Repšytė, J. Soft-sensors based on Black-box Models for Bioreactors Monitoring and State Estimation. In Proceedings of the 2020 12th International Conference on Bioinformatics and Biomedical Technology, Xi’an, China, 22–24 May 2020; pp. 43–47. [Google Scholar]
  48. Rösner, L.S.; Walter, F.; Ude, C.; John, G.T.; Beutel, S. Sensors and Techniques for On-Line Determination of Cell Viability in Bioprocess Monitoring. Bioengineering 2022, 9, 762. [Google Scholar] [CrossRef] [PubMed]
  49. Murugan, C. Soft Sensors for Biomass Monitoring during Low Cost Cellulase Production. In Biomass; IntechOpen: London, UK, 2021. [Google Scholar]
  50. Randek, J.; Mandenius, C.F. On-line soft sensing in upstream bioprocessing. Crit. Rev. Biotechnol. 2018, 38, 106–121. [Google Scholar] [CrossRef] [PubMed]
  51. Food and Drug Administration. Guidance for Industry, PAT-A Framework for Innovative Pharmaceutical Development, Manufacturing and Quality Assurance. 2004. Available online: https://www.fda.gov/media/71012/download (accessed on 1 February 2023).
  52. Tsopanoglou, A.; del Val, I.J. Moving towards an era of hybrid modelling: Advantages and challenges of coupling mechanistic and data-driven models for upstream pharmaceutical bioprocesses. Curr. Opin. Chem. Eng. 2021, 32, 100691. [Google Scholar] [CrossRef]
  53. Joiner, J.; Huang, Z.; McHugh, K.; Stebbins, M.; Aron, K.; Borys, M.; Khetan, A. Process modeling of recombinant adeno-associated virus production in HEK293 cells. Curr. Opin. Chem. Eng. 2022, 36, 100823. [Google Scholar] [CrossRef]
  54. Kang, W.; Wang, L.; Harrell, H.; Liu, J.; Thomas, D.L.; Mayfield, T.L.; Scotti, M.M.; Ye, G.J.; Veres, G.; Knop, D.R. An efficient rHSV-based complementation system for the production of multiple rAAV vector serotypes. Gene Ther. 2009, 16, 229–239. [Google Scholar] [CrossRef]
  55. Hartmann, J.; Münch, R.C.; Freiling, R.T.; Schneider, I.C.; Dreier, B.; Samukange, W.; Koch, J.; Seeger, M.A.; Plückthun, A.; Buchholz, C.J. A Library-Based Screening Strategy for the Identification of DARPins as Ligands for Receptor-Targeted AAV and Lentiviral Vectors. Mol. Ther.-Methods Clin. Dev. 2018, 10, 128–143. [Google Scholar] [CrossRef]
  56. Strobel, B.; Zuckschwerdt, K.; Zimmermann, G.; Mayer, C.; Eytner, R.; Rechtsteiner, P.; Kreuz, S.; Lamla, T. Standardized, Scalable, and Timely Flexible Adeno-Associated Virus Vector Production Using Frozen High-Density HEK-293 Cell Stocks and CELLdiscs. Hum. Gene Ther. Methods 2019, 30, 23–33. [Google Scholar] [CrossRef]
  57. Ghani, K.; Cottin, S.; Kamen, A.; Caruso, M. Generation of a high-titer packaging cell line for the production of retroviral vectors in suspension and serum-free media. Gene Ther. 2007, 14, 1705–1711. [Google Scholar] [CrossRef]
  58. Chhatre, S. Modelling approaches for bio-manufacturing operations. In Measurement, Monitoring, Modelling and Control of Bioprocesses; Springer: Berlin/Heidelberg, Germany, 2012; pp. 85–107. [Google Scholar]
  59. Udugama, I.A.; Lopez, P.C.; Gargalo, C.L.; Li, X.; Bayer, C.; Gernaey, K.V. Digital Twin in biomanufacturing: Challenges and opportunities towards its implementation. Syst. Microbiol. Biomanufacturing 2021, 1, 257–274. [Google Scholar] [CrossRef]
  60. Kotidis, P.; Pappas, I.; Avraamidou, S.; Pistikopoulos, E.N.; Kontoravdi, C.; Papathanasiou, M.M. DigiGlyc: A hybrid tool for reactive scheduling in cell culture systems. Comput. Chem. Eng. 2021, 154, 107460. [Google Scholar] [CrossRef]
  61. Fortuna, L.; Graziani, S.; Rizzo, A.; Xibilia, M.G. Soft Sensors for Monitoring and Control of Industrial Processes; Springer: Berlin/Heidelberg, Germany, 2007; Volume 22. [Google Scholar]
  62. Jiang, Y.; Yin, S.; Dong, J.; Kaynak, O. A review on soft sensors for monitoring, control and optimization of industrial processes. IEEE Sensors J. 2020, 21, 12868–12881. [Google Scholar] [CrossRef]
  63. Brunner, V.; Siegl, M.; Geier, D.; Becker, T. Challenges in the development of soft sensors for bioprocesses: A critical review. Front. Bioeng. Biotechnol. 2021, 9, 730. [Google Scholar] [CrossRef] [PubMed]
  64. Reyes, S.J.; Durocher, Y.; Pham, P.L.; Henry, O. Modern Sensor Tools and Techniques for Monitoring, Controlling, and Improving Cell Culture Processes. Processes 2022, 10, 189. [Google Scholar] [CrossRef]
  65. Shirsat, N.P.; English, N.J.; Glennon, B.; Al-Rubeai, M. Modelling of mammalian cell cultures. In Animal Cell Culture; Springer: Cham, Switzerland, 2015; pp. 259–326. [Google Scholar]
  66. Fernandes, R.L.; Bodla, V.K.; Carlquist, M.; Heins, A.L.; Lantz, A.E.; Sin, G.; Gernaey, K.V. Applying mechanistic models in bioprocess development. Meas. Monit. Model. Control. Bioprocesses 2012, 132, 137–166. [Google Scholar]
  67. Papathanasiou, M.M.; Kontoravdi, C. Engineering challenges in therapeutic protein product and process design. Curr. Opin. Chem. Eng. 2020, 27, 81–88. [Google Scholar] [CrossRef]
  68. Lopez, P.C.; Udugama, I.A.; Thomsen, S.T.; Roslander, C.; Junicke, H.; Mauricio-Iglesias, M.; Gernaey, K.V. Towards a digital twin: A hybrid data-driven and mechanistic digital shadow to forecast the evolution of lignocellulosic fermentation. Biofuels Bioprod. Biorefining 2020, 14, 1046–1060. [Google Scholar] [CrossRef]
  69. Sommeregger, W.; Sissolak, B.; Kandra, K.; von Stosch, M.; Mayer, M.; Striedner, G. Quality by control: Towards model predictive control of mammalian cell culture bioprocesses. Biotechnol. J. 2017, 12, 1600546. [Google Scholar] [CrossRef]
  70. Choi, H.K.; Son, S.H.; Sang-Il Kwon, J. Inferential model predictive control of continuous pulping under grade transition. Ind. Eng. Chem. Res. 2021, 60, 3699–3710. [Google Scholar] [CrossRef]
  71. Ohadi, S.K. Development of Soft Sensors for Monitoring of Chinese Hamster Ovary Cell Processes. 2014. Available online: http://hdl.handle.net/10012/8979 (accessed on 1 February 2023).
  72. Maruthamuthu, M.K.; Rudge, S.R.; Ardekani, A.M.; Ladisch, M.R.; Verma, M.S. Process analytical technologies and data analytics for the manufacture of monoclonal antibodies. Trends Biotechnol. 2020, 38, 1169–1186. [Google Scholar] [CrossRef]
  73. Arnold, S.A.; Crowley, J.; Woods, N.; Harvey, L.M.; McNeil, B. In-situ near infrared spectroscopy to monitor key analytes in mammalian cell cultivation. Biotechnol. Bioeng. 2003, 84, 13–19. [Google Scholar] [CrossRef] [PubMed]
  74. Rathore, A.S.; Thakur, G.; Nikita, S.; Banerjee, S. Control of Continuous Manufacturing Processes for Production of Monoclonal Antibodies. In Process Control, Intensification, and Digitalisation in Continuous Biomanufacturing; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2022; pp. 39–74. [Google Scholar]
  75. Zhang, A.H.; Zhu, K.Y.; Zhuang, X.Y.; Liao, L.X.; Huang, S.Y.; Yao, C.Y.; Fang, B.S. A robust soft sensor to monitor 1, 3-propanediol fermentation process by Clostridium butyricum based on artificial neural network. Biotechnol. Bioeng. 2020, 117, 3345–3355. [Google Scholar] [CrossRef] [PubMed]
  76. Brestrich, N.; Briskot, T.; Osberghaus, A.; Hubbuch, J. A tool for selective inline quantification of co-eluting proteins in chromatography using spectral analysis and partial least squares regression. Biotechnol. Bioeng. 2014, 111, 1365–1373. [Google Scholar] [CrossRef]
  77. Capito, F.; Skudas, R.; Kolmar, H.; Hunzinger, C. At-line mid infrared spectroscopy for monitoring downstream processing unit operations. Process. Biochem. 2015, 50, 997–1005. [Google Scholar] [CrossRef]
  78. Brunner, V.; Klöckner, L.; Kerpes, R.; Geier, D.U.; Becker, T. Online sensor validation in sensor networks for bioprocess monitoring using swarm intelligence. Anal. Bioanal. Chem. 2020, 412, 2165–2175. [Google Scholar] [CrossRef] [PubMed]
  79. Grigs, O.; Bolmanis, E.; Galvanauskas, V. Application of in-situ and soft-sensors for estimation of recombinant P. pastoris GS115 biomass concentration: A case analysis of HBcAg (Mut+) and HBsAg (MutS) production processes under varying conditions. Sensors 2021, 21, 1268. [Google Scholar] [CrossRef]
  80. Bayer, B.; von Stosch, M.; Melcher, M.; Duerkop, M.; Striedner, G. Soft sensor based on 2D-fluorescence and process data enabling real-time estimation of biomass in Escherichia coli cultivations. Eng. Life Sci. 2020, 20, 26–35. [Google Scholar] [CrossRef] [PubMed]
  81. Hrnčiřík, P.; Moucha, T.; Mareš, J.; Náhlík, J.; Janáčová, D. Software sensors for biomass concentration estimation in filamentous microorganism cultivation process. Chem. Biochem. Eng. Q. 2019, 33, 141–151. [Google Scholar] [CrossRef]
  82. Kornecki, M.; Strube, J. Process analytical technology for advanced process control in biologics manufacturing with the aid of macroscopic kinetic modeling. Bioengineering 2018, 5, 25. [Google Scholar] [CrossRef]
  83. Krishna, V.V.; Pappa, N.; Rani, S.J.V. Deep Learning based Soft Sensor for Bioprocess Application. In Proceedings of the 2021 IEEE Second International Conference on Control, Measurement and Instrumentation (CMI), Kolkata, India, 8–10 January 2021; IEEE: Piscataway, NJ, USA, 2021; pp. 155–159. [Google Scholar]
  84. Kornecki, M.; Strube, J. Accelerating biologics manufacturing by upstream process modelling. Processes 2019, 7, 166. [Google Scholar] [CrossRef]
  85. Solomatine, D.; See, L.M.; Abrahart, R. Data-driven modelling: Concepts, approaches and experiences. In Practical Hydroinformatics; Springer: Berlin/Heidelberg, Germany, 2009; pp. 17–30. [Google Scholar]
  86. Sun, Q.; Ge, Z. A survey on deep learning for data-driven soft sensors. IEEE Trans. Ind. Inform. 2021, 17, 5853–5866. [Google Scholar] [CrossRef]
  87. Kim, J.; Park, S.; Min, D.; Kim, W. Comprehensive Survey of Recent Drug Discovery Using Deep Learning. Int. J. Mol. Sci. 2021, 22, 9983. [Google Scholar] [CrossRef] [PubMed]
  88. Pra, M.D. Time Series Forecasting with Deep Learning and Attention Mechanism. 2020. Available online: https://towardsdatascience.com/time-series-forecasting-with-deep-learning-and-attention-mechanism-2d001fc871fc (accessed on 8 March 2021).
  89. Zhang, D.; Del Rio-Chanona, E.A.; Petsagkourakis, P.; Wagner, J. Hybrid physics-based and data-driven modeling for bioprocess online simulation and optimization. Biotechnol. Bioeng. 2019, 116, 2919–2930. [Google Scholar] [CrossRef] [PubMed]
  90. Emerson, J.; Kara, B.; Glassey, J. Multivariate data analysis in cell gene therapy manufacturing. Biotechnol. Adv. 2020, 45, 107637. [Google Scholar] [CrossRef] [PubMed]
  91. Solle, D.; Hitzmann, B.; Herwig, C.; Pereira Remelhe, M.; Ulonska, S.; Wuerth, L.; Prata, A.; Steckenreiter, T. Between the poles of data-driven and mechanistic modeling for process operation. Chem. Ing. Tech. 2017, 89, 542–561. [Google Scholar] [CrossRef]
  92. Iglesias, C.; Mehta, V.; Bolig, M. Handling Massive Proportion of Missing Labels in Multivariate Long-Term Time Series Forecasting. J. Phys. Conf. Ser. 2021, 2090, 012170. [Google Scholar] [CrossRef]
  93. Moser, A.; Appl, C.; Brüning, S.; Hass, V.C. Mechanistic mathematical models as a basis for digital twins. In Digital Twins; Springer: Berlin/Heidelberg, Germany, 2021; pp. 133–180. [Google Scholar]
  94. Dua, V.; Dua, P. A simultaneous approach for parameter estimation of a system of ordinary differential equations, using artificial neural network approximation. Ind. Eng. Chem. Res. 2012, 51, 1809–1814. [Google Scholar] [CrossRef]
  95. Kyriakopoulos, S.; Ang, K.S.; Lakshmanan, M.; Huang, Z.; Yoon, S.; Gunawan, R.; Lee, D.Y. Kinetic modeling of mammalian cell culture bioprocessing: The quest to advance biomanufacturing. Biotechnol. J. 2018, 13, 1700229. [Google Scholar] [CrossRef]
  96. Petre, E.; Selişteanu, D. Model approximation and simulations of a class of nonlinear propagation bioprocesses. In Numerical Analysis-Theory And Application; InTech: Rijeka, Croatia, 2011; pp. 211–230. [Google Scholar]
  97. Dürr, R.; Waldherr, S. Hybrid simulation algorithm for efficient numerical solution of population balance equations. IFAC-PapersOnLine 2018, 51, 290–295. [Google Scholar] [CrossRef]
  98. Xu, P. Analytical solution for a hybrid Logistic-Monod cell growth model in batch and continuous stirred tank reactor culture. Biotechnol. Bioeng. 2020, 117, 873–878. [Google Scholar] [CrossRef]
  99. Lee, K.; Parish, E. Parameterized neural ordinary differential equations: Applications to computational physics problems. Proc. R. Soc. A 2021, 477, 20210162. [Google Scholar] [CrossRef]
  100. Okereke, R.; Maliki, O.; Oruh, B. A novel method for solving ordinary differential equations with artificial neural networks. Appl. Math. 2021, 12, 900–918. [Google Scholar] [CrossRef]
  101. Yang, Y.; Hou, M.; Luo, J. A novel improved extreme learning machine algorithm in solving ordinary differential equations by Legendre neural network methods. Adv. Differ. Equ. 2018, 2018, 1–24. [Google Scholar] [CrossRef] [Green Version]
  102. Bradley, W.; Boukouvala, F. Two-Stage Approach to Parameter Estimation of Differential Equations Using Neural ODEs. Ind. Eng. Chem. Res. 2021, 60, 16330–16344. [Google Scholar] [CrossRef]
  103. Rackauckas, C.; Innes, M.; Ma, Y.; Bettencourt, J.; White, L.; Dixit, V. Diffeqflux. jl-A julia library for neural differential equations. arXiv 2019, arXiv:1902.02376. [Google Scholar]
  104. Xia, H.; Suliafu, V.; Ji, H.; Nguyen, T.; Bertozzi, A.; Osher, S.; Wang, B. Heavy ball neural ordinary differential equations. Adv. Neural Inf. Process. Syst. 2021, 34, 18646–18659. [Google Scholar]
  105. Chen, R.; Rubanova, Y.; Bettencourt, J.; Duvenaud, D. Neural ordinary differential equations. arXiv 2018, arXiv:1806.07366. [Google Scholar]
  106. Dhadphale, J.; Unni, V.; Saha, A.; Sujith, R. Neural ODE to model and prognose thermoacoustic instability. Chaos Interdiscip. J. Nonlinear Sci. 2022, 32, 013131. [Google Scholar] [CrossRef]
  107. Alahmadi, A.; Flegg, J.; Cochrane, D.; Drovandi, C.; Keith, J. A comparison of approximate versus exact techniques for Bayesian parameter inference in nonlinear ordinary differential equation models. R. Soc. Open Sci. 2020, 7, 191315. [Google Scholar] [CrossRef]
  108. Ge, H.; Xu, K.; Ghahramani, Z. Turing: A language for flexible probabilistic inference. In Proceedings of the International Conference on Artificial Intelligence And Statistics, Playa Blanca, Spain, 9–11 April 2018; pp. 1682–1690. [Google Scholar]
  109. Xing, Z.; Bishop, N.; Leister, K.; Li, Z.J. Modeling kinetics of a large-scale fed-batch CHO cell culture by Markov chain Monte Carlo method. Biotechnol. Prog. 2010, 26, 208–219. [Google Scholar] [CrossRef] [PubMed]
  110. Nguyen, T.N.; Sha, S.; Hong, M.S.; Maloney, A.J.; Barone, P.W.; Neufeld, C.; Wolfrum, J.; Springs, S.L.; Sinskey, A.J.; Braatz, R.D. Mechanistic model for production of recombinant adeno-associated virus via triple transfection of HEK293 cells. Mol. Ther.-Methods Clin. Dev. 2021, 21, 642–655. [Google Scholar] [CrossRef] [PubMed]
  111. Iglesias, C.F., Jr.; Xu, X.; Mehta, V.; Akassou, M.; Venereo-Sanchez, A.; Belacel, N.; Kamen, A.; Bolic, M. Monitoring the Recombinant Adeno-Associated Virus Production using Extended Kalman Filter. Processes 2022, 10, 2180. [Google Scholar] [CrossRef]
  112. Von Stosch, M.; Oliveira, R.; Peres, J.; de Azevedo, S.F. Hybrid semi-parametric modeling in process systems engineering: Past, present and future. Comput. Chem. Eng. 2014, 60, 86–101. [Google Scholar] [CrossRef]
  113. Bayer, B.; Striedner, G.; Duerkop, M. Hybrid modeling and intensified DoE: An approach to accelerate upstream process characterization. Biotechnol. J. 2020, 15, 2000121. [Google Scholar] [CrossRef] [PubMed]
  114. Sokolov, M.; von Stosch, M.; Narayanan, H.; Feidl, F.; Butté, A. Hybrid modeling—A key enabler towards realizing digital twins in biopharma? Curr. Opin. Chem. Eng. 2021, 34, 100715. [Google Scholar] [CrossRef]
  115. Pinto, J.; Mestre, M.; Ramos, J.; Costa, R.S.; Striedner, G.; Oliveira, R. A general deep hybrid model for bioreactor systems: Combining first principles with deep neural networks. Comput. Chem. Eng. 2022, 165, 107952. [Google Scholar] [CrossRef]
  116. Luo, Y.; Kurian, V.; Ogunnaike, B.A. Bioprocess systems analysis, modeling, estimation, and control. Curr. Opin. Chem. Eng. 2021, 33, 100705. [Google Scholar] [CrossRef]
  117. Gimpel, A.L.; Katsikis, G.; Sha, S.; Maloney, A.J.; Hong, M.S.; Nguyen, T.N.; Wolfrum, J.; Springs, S.L.; Sinskey, A.J.; Manalis, S.R.; et al. Analytical methods for process and product characterization of recombinant adeno-associated virus-based gene therapies. Mol. Ther.-Methods Clin. Dev. 2021, 20, 740–754. [Google Scholar] [CrossRef]
  118. Ohadi, K.; Aghamohseni, H.; Legge, R.L.; Budman, H.M. Fluorescence-based soft sensor for at situ monitoring of chinese hamster ovary cell cultures. Biotechnol. Bioeng. 2014, 111, 1577–1586. [Google Scholar] [CrossRef]
  119. Claßen, J.; Graf, A.; Aupert, F.; Solle, D.; Höhse, M.; Scheper, T. A novel LED-based 2D-fluorescence spectroscopy system for in-line bioprocess monitoring of Chinese hamster ovary cell cultivations—Part II. Eng. Life Sci. 2019, 19, 341–351. [Google Scholar] [CrossRef]
  120. Whelan, J.; Craven, S.; Glennon, B. In situ Raman spectroscopy for simultaneous monitoring of multiple process parameters in mammalian cell culture bioreactors. Biotechnol. Prog. 2012, 28, 1355–1362. [Google Scholar] [CrossRef] [PubMed]
  121. Abu-Absi, N.R.; Kenty, B.M.; Cuellar, M.E.; Borys, M.C.; Sakhamuri, S.; Strachan, D.J.; Hausladen, M.C.; Li, Z.J. Real time monitoring of multiple parameters in mammalian cell culture bioreactors using an in-line Raman spectroscopy probe. Biotechnol. Bioeng. 2011, 108, 1215–1221. [Google Scholar] [CrossRef] [PubMed]
  122. Berry, B.; Moretto, J.; Matthews, T.; Smelko, J.; Wiltberger, K. Cross-scale predictive modeling of CHO cell culture growth and metabolites using R aman spectroscopy and multivariate analysis. Biotechnol. Prog. 2015, 31, 566–577. [Google Scholar] [CrossRef]
  123. Faassen, S.M.; Hitzmann, B. Fluorescence spectroscopy and chemometric modeling for bioprocess monitoring. Sensors 2015, 15, 10271–10291. [Google Scholar] [CrossRef]
  124. Brownlee, J. Deep Learning for Time Series Forecasting: Predict the Future with MLPs, CNNs and LSTMs in Python; Machine Learning Mastery: San Juan, Puerto Rico, 2018. [Google Scholar]
  125. Brownlee, J. Long Short-Term Memory Networks with Python: Develop Sequence Prediction Models with Deep Learning; Machine Learning Mastery: San Juan, Puerto Rico, 2017. [Google Scholar]
  126. Lim, B.; Zohren, S. Time-series forecasting with deep learning: A survey. Philos. Trans. R. Soc. A 2021, 379, 20200209. [Google Scholar] [CrossRef] [PubMed]
  127. Torres, J.F.; Hadjout, D.; Sebaa, A.; Martínez-Álvarez, F.; Troncoso, A. Deep learning for time series forecasting: A survey. Big Data 2021, 9, 3–21. [Google Scholar] [CrossRef]
  128. Curreri, F.; Patanè, L.; Xibilia, M.G. RNN-and LSTM-Based Soft Sensors Transferability for an Industrial Process. Sensors 2021, 21, 823. [Google Scholar] [CrossRef] [PubMed]
  129. Li, D.; Li, Z.; Sun, K. Development of a novel soft sensor with long short-term memory network and normalized mutual information feature selection. Math. Probl. Eng. 2020, 2020, 7617010. [Google Scholar] [CrossRef]
  130. Shen, F.; Zheng, J.; Ye, L.; Ma, X. LSTM Soft Sensor Development of Batch Processes With Multivariate Trajectory-Based Ensemble Just-in-Time Learning. IEEE Access 2020, 8, 73855–73864. [Google Scholar] [CrossRef]
  131. Vaněk, M.; Hrnčiřík, P.; Vovsik, J.; Nahlik, J. On-line estimation of biomass concentration using a neural network and information about metabolic state. Bioprocess Biosyst. Eng. 2004, 27, 9–15. [Google Scholar] [CrossRef]
  132. Singh, S.; Rani, A.; Singh, V.; Yadav, J. Soft sensor for inferential control in non-isothermal CSTR. In Proceedings of the 2016 International Conference on Advances in Computing, Communications and Informatics (ICACCI), Jaipur, India, 21–24 September 2016; IEEE: Piscataway, NJ, USA, 2016; pp. 1785–1790. [Google Scholar]
  133. Mandenius, C.F.; Gustavsson, R. Mini-review: Soft sensors as means for PAT in the manufacture of bio-therapeutics. J. Chem. Technol. Biotechnol. 2015, 90, 215–227. [Google Scholar] [CrossRef]
  134. Zhu, X.; Rehman, K.U.; Wang, B.; Shahzad, M. Modern soft-sensing modeling methods for fermentation processes. Sensors 2020, 20, 1771. [Google Scholar] [CrossRef] [PubMed]
  135. Kiviharju, K.; Salonen, K.; Moilanen, U.; Eerikäinen, T. Biomass measurement online: The performance of in situ measurements and software sensors. J. Ind. Microbiol. Biotechnol. 2008, 35, 657–665. [Google Scholar] [CrossRef] [PubMed]
  136. Yuan, X.; Li, L.; Wang, Y. Nonlinear dynamic soft sensor modeling with supervised long short-term memory network. IEEE Trans. Ind. Inform. 2019, 16, 3168–3176. [Google Scholar] [CrossRef]
  137. Pisa, I.; Santin, I.; Morell, A.; Vicario, J.L.; Vilanova, R. LSTM-Based Wastewater Treatment Plants Operation Strategies for Effluent Quality Improvement. IEEE Access 2019, 7, 159773–159786. [Google Scholar] [CrossRef]
  138. Gawlikowski, J.; Tassi, C.R.N.; Ali, M.; Lee, J.; Humt, M.; Feng, J.; Kruspe, A.; Triebel, R.; Jung, P.; Roscher, R.; et al. A survey of uncertainty in deep neural networks. arXiv 2021, arXiv:2107.03342. [Google Scholar]
  139. Bayer, B.; von Stosch, M.; Striedner, G.; Duerkop, M. Comparison of modeling methods for DoE-based holistic upstream process characterization. Biotechnol. J. 2020, 15, 1900551. [Google Scholar] [CrossRef] [Green Version]
  140. Kozma, B.; Salgó, A.; Gergely, S. On-line glucose monitoring by near infrared spectroscopy during the scale up steps of mammalian cell cultivation process development. Bioprocess Biosyst. Eng. 2019, 42, 921–932. [Google Scholar] [CrossRef]
  141. Tamburini, E.; Marchetti, M.G.; Pedrini, P. Monitoring key parameters in bioprocesses using near-infrared technology. Sensors 2014, 14, 18941–18959. [Google Scholar] [CrossRef]
  142. Zimmerleiter, R.; Kager, J.; Nikzad-Langerodi, R.; Berezhinskiy, V.; Westad, F.; Herwig, C.; Brandstetter, M. Probeless non-invasive near-infrared spectroscopic bioprocess monitoring using microspectrometer technology. Anal. Bioanal. Chem. 2020, 412, 2103–2109. [Google Scholar] [CrossRef]
  143. Reardon, K.F. Practical monitoring technologies for cells and substrates in biomanufacturing. Curr. Opin. Biotechnol. 2021, 71, 225–230. [Google Scholar] [CrossRef]
  144. Jin, S.; Sun, F.; Hu, Z.; Li, Y.; Zhao, Z.; Du, G.; Shi, G.; Chen, J. Online quantitative substrate, product, and cell concentration in citric acid fermentation using near-infrared spectroscopy combined with chemometrics. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 285, 121842. [Google Scholar] [CrossRef] [PubMed]
  145. Gomes, J.; Chopda, V.; Rathore, A.S. Monitoring and control of bioreactor: Basic concepts and recent advances. In Bioprocessing Technology for Production of Biopharmaceuticals and Bioproducts; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2018; pp. 201–237. [Google Scholar]
  146. Gerzon, G.; Sheng, Y.; Kirkitadze, M. Process analytical technologies–advances in bioprocess integration and future perspectives. J. Pharm. Biomed. Anal. 2022, 207, 114379. [Google Scholar] [CrossRef] [PubMed]
  147. Sharma, N.; Liu, Y. A hybrid science-guided machine learning approach for modeling chemical processes: A review. AIChE J. 2022, 68, e17609. [Google Scholar] [CrossRef]
  148. Selişteanu, D.; Șendrescu, D.; Georgeanu, V.; Roman, M. Mammalian cell culture process for monoclonal antibody production: Nonlinear modelling and parameter estimation. BioMed Res. Int. 2015, 2015, 598721. [Google Scholar] [CrossRef] [PubMed]
  149. Kourti, T. Multivariate statistical process control and process control using latent variables. Compr. Chemom. 2020, 275–303. Available online: https://www.sciencedirect.com/science/article/pii/B9780124095472148875 (accessed on 1 February 2023).
  150. Simutis, R.; Jenzsch, M.; Lübbert, A. Model-Based Technologies Enabling Optimal Bioreactor Performance. In Bioprocessing Technology for Production of Biopharmaceuticals and Bioproducts; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2018; pp. 175–200. [Google Scholar]
  151. Narayanan, H.; Behle, L.; Luna, M.F.; Sokolov, M.; Guillén-Gosálbez, G.; Morbidelli, M.; Butté, A. Hybrid-EKF: Hybrid model coupled with extended Kalman filter for real-time monitoring and control of mammalian cell culture. Biotechnol. Bioeng. 2020, 117, 2703–2714. [Google Scholar] [CrossRef]
  152. Simutis, R.; Lübbert, A. Hybrid approach to state estimation for bioprocess control. Bioengineering 2017, 4, 21. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Soft sensor development approaches based on DDM for rAAV production monitoring. A possible approach is use unstructured MM to generate an information-rich training dataset and DDM for multi-step forecasting of rAAV viral titer.
Figure 1. Soft sensor development approaches based on DDM for rAAV production monitoring. A possible approach is use unstructured MM to generate an information-rich training dataset and DDM for multi-step forecasting of rAAV viral titer.
Bioengineering 10 00229 g001
Figure 2. Soft sensor development approach based on HM for rAAV production monitoring. A possible approach is combining unstructured MM, DDM and EKF for real-time monitoring of rAAV viral titer.
Figure 2. Soft sensor development approach based on HM for rAAV production monitoring. A possible approach is combining unstructured MM, DDM and EKF for real-time monitoring of rAAV viral titer.
Bioengineering 10 00229 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Iglesias, C.F., Jr.; Ristovski, M.; Bolic, M.; Cuperlovic-Culf, M. rAAV Manufacturing: The Challenges of Soft Sensing during Upstream Processing. Bioengineering 2023, 10, 229. https://doi.org/10.3390/bioengineering10020229

AMA Style

Iglesias CF Jr., Ristovski M, Bolic M, Cuperlovic-Culf M. rAAV Manufacturing: The Challenges of Soft Sensing during Upstream Processing. Bioengineering. 2023; 10(2):229. https://doi.org/10.3390/bioengineering10020229

Chicago/Turabian Style

Iglesias, Cristovão Freitas, Jr., Milica Ristovski, Miodrag Bolic, and Miroslava Cuperlovic-Culf. 2023. "rAAV Manufacturing: The Challenges of Soft Sensing during Upstream Processing" Bioengineering 10, no. 2: 229. https://doi.org/10.3390/bioengineering10020229

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop