Next Article in Journal
Investigating the Mechanical Behaviour of Viscoelastic and Brittle Pharmaceutical Excipients During Tabletting: Revealing the Unobvious Potential of Advanced Compaction Simulation
Previous Article in Journal
Pharmacokinetic Landscape and Interaction Potential of SGLT2 Inhibitors: Bridging In Vitro Findings and Clinical Implications
Previous Article in Special Issue
Improved Dissolution of Poorly Water-Soluble Rutin via Solid Dispersion Prepared Using a Fluid-Bed Coating System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments in Pharmaceutical Spray Drying: Modeling, Process Optimization, and Emerging Trends with Machine Learning

Department of Chemical and Petroleum Engineering, United Arab Emirates University, Sheikh Khalifa Bin Zayed Street, Al-Ain 15551, United Arab Emirates
*
Author to whom correspondence should be addressed.
Pharmaceutics 2025, 17(12), 1605; https://doi.org/10.3390/pharmaceutics17121605
Submission received: 1 August 2025 / Revised: 17 October 2025 / Accepted: 21 October 2025 / Published: 13 December 2025

Abstract

Spray drying techniques are widely used in the pharmaceutical industry to produce fine drug powders with different properties depending on the route of administration. Process parameters play a vital role in the critical quality attributes of the final product. This review highlights the progress and challenges in modeling the spray-drying process, with a focus on pharmaceutical applications. Computational fluid dynamics (CFD) is a well-known method used for the modeling and numerical simulation of spray drying processes. However, owing to their limitations, including high computational costs, experimental validation, and limited accuracy under complex spray drying conditions. Machine learning (ML) models have recently emerged as integral tools for modeling/optimizing the spray drying process. Despite promising accuracy, ML models depend on high-quality data and may fail to predict the influence of new formulation or process parameters on the properties of the dried powder. This review outlines the shortcomings of CFD modeling in the spray drying process. A hybrid model combining ML and CFD models, emerging techniques such as the digital twin approach, transfer learning, and explainable AI (XAI) are also discussed. A hybrid model combining ML and CFD models is also discussed. ML is considered an emerging technique that could assist the spray drying process, and most importantly, the utilization of this method in pharmaceutical spray drying.

Graphical Abstract

1. Introduction

Spray drying is an energy-intensive, fast, continuous, and highly efficient method for producing solid particles from liquid solutions, suspensions, emulsions, slurries, pastes, and melts [1]. This process can produce nano- to micron-sized particles in a short time, ca. 2–4 s. Spray drying has various applications in the food, chemical, biochemical, and pharmaceutical industries [2]. Spray drying is frequently used in the food industry to produce various food products, such as milk powder, cheese powder, egg powder, juice powder, instant coffee/tea powder, color, and flavoring ingredients [3]. Fruit and vegetable powders produced by spray drying are more stable than liquid juice and can reduce the costs associated with packaging, storage, and transportation [4]. Moreover, spray-dried fruit and vegetable powders can be preserved for a longer period, depending on the packaging, compared to their liquid form [5]. Spray drying is widely used to encapsulate essential oils [6] and probiotic powders in dairy products [7]. The spray drying market is huge and growing worldwide, with market share estimated to reach between $6–$7 billion by 2025 [8]. The food industry is the biggest market for spray drying equipment, with a compound annual growth rate (CAGR) varying from 5.7% to 6.8% between 2019 and 2025 [8,9] The pharmaceutical sector is responsible for over 80% of the market’s value in terms of products such as active pharmaceutical ingredients (APIs) and excipients. It is expected to achieve a steady 17% growth rate until 2028 [10].
The spray drying process has various benefits, such as ease of operation, energy efficiency, and low operational costs; control over particle size, shape, and morphology; high encapsulation efficiency and product stability; and the ability to be scaled to the industrial level [11]. Some of the other currently available methods for drying of products include freeze drying, fluidized bed drying, spray–freeze drying, electro spraying, solar drying, superheated-steam drying, infrared drying, and supercritical fluid drying. Although these methods have been widely used for drying particles in various applications, such as food, pharmaceuticals, and chemicals, there are many limitations to these techniques, as summarized in Table 1.
In the pharmaceutical industry, spray drying is frequently used to alter the size distribution, crystallinity, morphology, and moisture content of spray-dried particles and to improve powder compactability [43,44,45]. Dry particles with narrow and respirable size distributions of 1–5 µm can be formulated using a spray drying process for effective pulmonary drug delivery [46]. Most modern therapeutic drugs show poor aqueous solubility and low drug dissolution rates, which reduce their bioavailability and slow the absorption process. Spray drying can produce a stable amorphous dispersion containing an active pharmaceutical ingredient (API) and polymer, thereby improving the drug dissolution rate [47]. Engineered particles produced by spray drying showed consistent dosing and deep lung deposition [48]. Spray drying permits the drying of heat-sensitive materials such as proteins or enzymes without losing their pharmacological benefits [49].
As illustrated in Figure 1, the spray drying process begins with the liquid feed solution being fed into the drying chamber along with hot gas, which atomizes the feed solution into fine droplets. The drying of droplets into fine powder occurs immediately as the hot drying gas simultaneously enters the drying chamber, either as a co-current or counter-current flow, and encounters the droplets [50]. The dried powder is then separated from the drying gas using a cyclone separator and collected in a collection vessel.
Spray-dried API powder generally consists of an excipient or carrier that acts as a binder, sweetener, or coating agent, or improves the performance of the API for different routes of drug delivery. Excipients such as cyclodextrin, poloxamers [50], and polydimethylsiloxanes [52] have been widely used for oral, nasal, and transdermal drug delivery. Drug particles less than 5 µm in size stick to larger carrier particles such as lactose (60–90 µm) through physical forces such as van der Waals, electrostatic, capillary, and mechanical interlocking forces that are vital for drug detachment and lung deposition during inhalation. The strength of these forces directly influences the efficiency of drug delivery from DPIs. Stronger forces can decrease dispersion and lung deposition [53] Molecular-level incompatibilities can happen between drug and an excipient, which can create crystal strain, influencing the uniformity and flowability of the powder mixture. The Maillard reaction is a specific chemical incompatibility that can result in dose reduction or toxic byproducts in DPI formulations [54]. Surface modification of excipients with magnesium stearate or Aerosil R972 can reduce these interactions, leading to improved bulk properties and DPI performance [55]. Spray-dried dry powders are generally delivered via dry powder inhalers (DPIs). Excipients such as lactose monohydrate have been used as coarse carriers in low-dose DPIs because of their nontoxicity, stability, and compatibility with most active drugs [56,57]. In high-dose DPIs, coarse carriers decrease the final powder volume by acting as a bulking agent. The major challenge of the micronized drug particles in high-dose DPIs is poor flowability and cohesion, resulting in particle agglomeration and incorrect dosing [58]. The addition of coarse carriers to the drug increases the bulk density, thereby improving the flowability of powder and reducing the dosage requirement compared to the drug alone [59]. Trehalose is another excipient that has been used as a carrier in DPI formulations, such as albuterol sulfate, disodium cromoglycate, and fluticasone propionate [60]. Amino acids such as L-leucine and tri-leucine are excipients that can improve the aerosol performance of spray-dried DPIs [61] and protect spray-dried inhalable powders from moisture [62]. D-mannitol is widely used as an excipient in pharmaceutical drugs because of its ability to absorb moisture, chemically inactive nature, and non-decomposable and non-degradable tendencies in reaction with drugs [63].
Nasal powders containing spray-dried excipients and drugs have demonstrated effective intranasal drug delivery in rabbits [64]. Bioavailability is one of the main concerns in oral drug delivery. Carrier particles, when spray-dried along with a poorly water-soluble drug, significantly improved oral bioavailability and drug efficiency in rats and mice compared to the pure drug [65]. Drug dissolution rate and fine particle fraction are vital for the bioavailability and therapeutic effectiveness of drugs. Carrier microparticles containing drug nanoparticles prepared by spray drying showed improved drug dissolution rates and higher fine-particle fractions for oral delivery [66]. For transdermal drug delivery, the spray-dried excipient drug powder decreased toxicity and improved the efficacy of treatment [67], whereas it showed the capability for parenteral drug delivery [68]. Spray-dried water-insoluble drug nanoparticles incorporated into water-soluble carrier particles improved drug absorption in rats and could be delivered via oral, pulmonary, and injection pathways [69].
The administration of pharmaceutical drugs by inhalation allows for more efficient absorption of drug molecules from the lungs, in contrast to oral, nasal, or transdermal routes [70]. Moreover, inhalation allows direct drug delivery to the target organs with minimal side effects [71]. DPIs are considered the most suitable for inhalation purposes because they are portable, economical, easy to handle, and environmentally friendly compared to nebulizers or pressurized metered dose inhalers and ensure better formulation stability than liquid dosage forms [72]. DPIs contain an active pharmaceutical ingredient of suitable aerodynamic size of 1–5 µm for inhalation [73]. Studies have suggested that the effectiveness of using spray-dried powder for the inhalation of spray-dried excipient powder enhances the clearance of mucus in the lungs [74,75]. Several researchers have suggested that for children affected by cystic fibrosis, inhaled dry powder shows remarkable improvements in lung function, regardless of age or disease severity [76,77,78,79]. Spray-dried powder showed bronchial responsiveness upon inhalation in people with asthma [80]. Spray-dried powder was found to be a suitable carrier in powder inhalers and can be produced with the desired properties by spray drying to meet the needs of the chosen drug [81]. The proportion of disintegrated nanoparticles increased with increasing excipient concentration, suggesting that a suitable excipient can convert nanoparticles into a dry inhalable powder by spray drying [82].
Spray-dried chitosan/mannitol/leucine–quercetin formulation for inhalation showed high yield (approximately 65%), encapsulation efficiency (>45%), and drug loading (12–26%) [83]. The drying temperature influenced the crystallization of the spray-dried excipients. Lower inlet temperatures (70 °C and 90 °C) lead to a slow release of moisture content and a longer crystallization period [84]. The increased inlet temperature negatively affected the aerodynamic performance of the spray-dried carrier-based formulations [85]. The particle morphology of the spray-dried inhalable carrier and protein molecules changed from spherical to irregular at high inlet temperatures, and more particle agglomerates were present in the final dry powder [86]. The yield of the excipient dry powder obtained by spray drying the aqueous excipient solution was slightly higher at lower inlet temperatures [87]. When spray-dried as a single solution and mixed with a drug, the binary carriers showed better aerosol performance than the single carriers [88]. Studies have indicated that the particle size and aerodynamic diameter of large spray-dried porous carrier particles are affected by changes in the inlet temperature, feed rate, and airflow rate. The particle size of the excipients is considered an important characteristic affecting tablet dissolution. The results showed that the specific surface area was the key factor affecting the dissolution of drug molecules produced by similar particle-size excipient grades [89]. Various amounts of mannitol were dissolved in a cyclosporine A/ethanol suspension and spray-dried at an inlet temperature of 120 °C, with outlet temperatures ranging from 46 to 59 °C, and an aspiration rate of 100%, an atomization flow rate set at 819 NL/h, and a liquid feed rate of 3.5 mL/min to produce respirable particles. The addition of mannitol to cyclosporine A improved the drug dissolution rate without affecting the particle size distribution, surface roughness, and aerosol performance [90]. Powder flowability is an important characteristic of pharmaceutical drugs that ensures a uniform tablet weight and produces tablets with consistent and reproducible features. Dry powders containing protein molecules and carrier particles prepared by spray drying exhibited a high fine particle fraction [91].
Excipients consisting of amorphous drugs show faster drug release rates than microcrystalline excipients [92]. The excipient concentration influenced the physical stability and aerosolization performance of the spray-dried protein powder. High excipient concentrations led to excipient crystallization and phase separation during storage, which negatively affected the physical stability and aerosolization performance of the spray-dried protein powder for inhalation [93].
Mathematical modeling and computational fluid dynamics offer enormous potential for optimizing spray drying process conditions, thereby making it economical and accelerating the experimental process [94,95,96].
A design of experiment (DoE) approach is often used to screen, optimize, and identify the critical and non-critical process parameters affecting the final powder properties, and several studies [97,98,99] have utilized different DoE approaches for spray drying.
At the industry level, design of experiments and other multivariate statistical approaches are commonly used to recognize critical process parameters (CPPs) and their relationship with product yield and critical quality attributes (CQAs). However, these models cannot effectively predict the effects of critical process parameters on the critical quality attributes [100]. Response surface methodology (RSM) combined with the DoE approach is being effectively utilized to optimize the process parameters required to produce pharmaceutical drugs. [101,102] However, RSM has a few limitations for nonlinear processes, such as spray drying [103]. The drawback of this approach is that it does not consider the effects of independent factors on the response variables. The model does not account for the interaction between independent variables, which might significantly affect the powder properties. A nonlinear model, such as machine learning (ML), can overcome this problem by removing the errors associated with DoE–based processes [104]. ML models illustrate the relationship between input variables and response parameters in a spray drying process by considering all the bias, systematic, and data errors for better accuracy [105]. These models can identify the optimum operating parameters, predict how variation in spray drying conditions affects the CQAs of the powder, and help in real-time monitoring of important parameters to maintain optimal drying conditions [106]. In a spray drying process, nozzle blockage or overheating can significantly affect the product quality. ML models can predict nozzle blockage to prevent interruptions during experiments and improve the process efficiency [106]. They are also used for predicting drying times, reducing energy consumption, formulation development, and developing advanced drug delivery systems [107]. However, the ML models also have a few limitations that need to be addressed. They require huge, high-quality data sets for predicting the powder properties [106]. Moreover, their effectiveness is limited by their complexity, generalizability [108], inability to evaluate new parameters, and negligence to process uncertainties [109].
The motivation behind this work was to review past and current modeling techniques for modeling the spray drying process. In the first part of the study, a general outline of the spray drying process is presented, including the transport phenomena occurring during the spray drying process and the influence of the process parameters on the properties of the final dried powder. The mathematical and computational fluid dynamics (CFD) modeling of the spray drying process, its limitations, and recommendations are highlighted in sec. 4. Then, the fundamentals of machine learning (ML) models, including the framework, hybrid ML model, introduction of relevant software, and analysis techniques, are presented in Section 5, followed by conclusions.

2. Transport Phenomena in Spray Drying

The spray drying process begins with the atomization of the liquid feed solution into droplets, owing to a decrease in surface tension. In the atomization process, the pressurized gas enters the nozzle and breaks down the liquid into small droplets, leading to an increase in their surface area, thereby enhancing the heat and mass transfer between the hot drying gas and droplets, as depicted in Figure 2A. The size and movement of the liquid droplets through the drying chamber and their interaction with the drying gas affect the drying. Small and large droplets in any type of atomizer spray dry at different rates owing to the different periods of capillary transport, diffusion, crystallization, and heat transport. In addition, different trajectories of small and large droplets evolve owing to different drag and inertial forces, and different exposures of small and large particles to the drying gas [110]. The flow of the air-droplets inside the drying chamber can be counter-current, co-current, or mixed-flow. The co-current flow type is mostly used in the pharmaceutical industry [111].

2.1. Droplet Drying Process

During spray drying, the droplets encounter hot drying air; they initially heat without significant evaporation (sensible heating), as shown in Figure 2B. During this stage, the temperature increases steadily until it reaches the wet-bulb temperature [112]. Additionally, it is assumed that droplet drying occurs at a constant evaporation rate (K) and that the droplet diameter decreases linearly from the initial diameter (d0). This period is known as the constant drying rate period and can be described by the d2-law, τ D = d 0 2 / K [113]. In the spray drying process, droplet diameters are generally less than 100 µm. The value of the corresponding Biot number is normally less than 0.1 [114]. Biot number is the ratio of thermal resistance for conduction inside the droplet to the resistance for convection at the droplet surface. If the Biot number is less than 0.1, the lumped heat capacity of the droplet can be assumed, indicating that the droplet temperature is uniformly distributed. The drying temperature, humidity, and velocity of the air surrounding the droplet surface are factors that should be considered during solvent evaporation from the atomized droplets [115,116,117,118].

2.2. Particle Formation Process

The transition from a constant-rate period to a falling-rate period occurs due to the increase in solid concentration on the droplet surface. At a critical moisture concentration, the evaporation rate decreases, indicating the beginning of the falling rate period. Solute diffusion occurs inside the droplet owing to the concentration gradient between the droplet surface and the core [119]. The continuous evaporation of the solvent resulted in droplet shrinkage, and a shell began to form around the droplet surface. The vapor diffuses through the formed shell, and the particle size becomes constant. The shell significantly reduced the mass transfer of the solvent to the droplet surface. The heat transfer to the droplet during this stage increases the particle temperature. Finally, the wet particles were completely transformed into dry solid particles. A schematic of the particle formation process is shown in Figure 3A.
The morphology of the formed particles can be related to the process operating parameters using a dimensionless number known as the Peclet number (Pe), which is defined as the ratio of the evaporation rate (K) to the diffusion coefficient (De). As outlined in Figure 3B, a lower Peclet number (Pe < 1) mostly leads to the formation of spherical solid particles, whereas a higher Pe value (Pe > 1) results in folded particles [120]. A high value of Peclet number, i.e., high drying temperature, results in a higher evaporation rate compared to the diffusion of primary particles in the droplet, which yields hollow particles. If the shell cannot withstand the external pressure, a folded particle is formed [121,122]. The droplet evaporation rate also influences the final particle morphology. A slow evaporation rate leads to denser particles, whereas a high evaporation rate causes skin formation on the droplet surface and produces less dense particles. Particle morphology can be explained in terms of size, shape, internal structure, and surface properties [123,124,125,126]. However, particle engineering requires a thorough knowledge of the particle formation process [127]. Therefore, it is necessary to understand and control particle formation to develop effective drug administration and particle engineering methods [127,128,129,130]. Moreover, artificial intelligence and machine learning (AI/ML) can play a significant role in bridging the influence of process parameters on the final particle morphology. ML models can predict spray-dried API particle size with errors between 7.7–18.6% [131]. Bayesian optimization can be utilized to achieve defined morphology while considering product properties such as yield, moisture, and FPF. Feature-importance/SHAP analyses indicate which input parameters (e.g., atomizing air vs. feed solids) dominantly control specific morphological outcomes, providing an advanced particle engineering hybrid method to tailor the final particle morphology of the dried powder [132,133].

3. Process Parameters

Several process parameters, such as inlet temperature, outlet temperature, feed concentration, feed rate, drying gas rate, and spray gas flow rate, affect the final properties of the spray-dried powder. The inlet temperature significantly affects the properties and quality of the final powder. A higher inlet temperature increases the powder yield and outlet temperatures [134,135]. Moreover, a high inlet temperature influences the particle formation process owing to the high rate of solvent evaporation. A higher inlet temperature leads to the rapid formation of a solid skin on the outer surface of the droplet that captures the solvent vapors. Faster drying rates, achieved through higher inlet temperatures, result in particles with a higher Tg (glass transition temperature), making them more suitable for product stability. In contrast, lower inlet temperatures result in low Tg products with stickier particles, which reduce the final yield due to wall deposition. Inlet temperature also influences the powder properties, such as particle size, moisture content (see Figure 4A), bulk density, and solubility. The moisture gradient generated inside the droplet due to high inlet temperature can influence the particle formation process, which in turn can affect the morphology of the dried powder [136]. Variation in the inlet temperature can lead to particles of different shapes and surface roughness’s [137,138,139]. Higher inlet air temperature resulted in smoother particles and increased the breaking strength of spray-dried mannitol powder for inhalation [140]. The inlet temperature also influences the breaking strength and crystallinity of the spray-dried powders. It has been reported that a higher inlet temperature resulted in a higher degree of crystallinity for spray-dried powder [141]. The optimal inlet temperature varies depending on the sample being dried and it allows appropriate solvent evaporation and prevents product degradation [142,143].
The outlet temperature (Tout) is a function of the feed flow rate, feed concentration, drying gas flow rate, and inlet temperature [118]. Variation in the outlet temperature did not influence the particle size, but it affected the morphology and crystallinity of the spray-dried powder, as shown in Figure 5 [146]. Reports show that when Tout varied between 90 °C and 157 °C, it resulted in a higher powder crystallinity [141]. Moreover, Tout should not be higher than the product glass transition temperature, because it increases the solid-phase crystallization rate [147]. Tout can also influence the residual solvent content. During droplet formation, film-producing polymers rapidly form a skin around the droplet surface, resulting in solvent entrapment, which makes solvent removal difficult [127]. Therefore, Tout should be maintained properly to remove the solvent content without affecting the solid dispersion stability. It has been proven that low Tout (60–80 °C) is needed to prevent excessive denaturation of protein [148], whereas higher Tout improved the storage stability of spray-dried probiotic dry powder [149].
Feed concentration can influence powder properties such as particle size, particle morphology, moisture content, and yield. It was reported that higher feed concentrations of aqueous mannitol solution resulted in particles with a rougher surface compared to lower feed concentrations. Moreover, higher feed concentrations lead to hollow particles with high porosities and low bulk densities, whereas lower feed concentrations produce smaller particles [140]. A higher feed concentration implies a lower solvent concentration in each droplet, resulting in high Pe values, shorter evaporation times, and porous, less dense particles in the final drying stage. Therefore, it increases the probability of agglomeration, leading to more porous particles with low density and rougher surfaces [151]. Also, increasing the feed concentration reduced the overall thermal efficiency, evaporative efficiency, material loss in the cyclone, powder moisture content, and bulk density, whereas increasing the particle size [152,153] and the yield of the powder.
Powder properties such as solvent evaporation rate, morphology, particle size, and density can be altered by varying the feed rate [94]. The feed flow rate is crucial because it influences the droplet size and relative humidity [154]. Higher moisture content was observed in the final dry powder at higher feed rates, resulting in a lower powder yield [151]. The effects of different feed flow rates on the moisture content of the spray-dried powders are shown in Figure 4B. Higher yields were obtained at lower feed rates. Moreover, higher feed flow rates increased the atomization of the liquid feed solution in the drying chamber and generated larger liquid droplets. As a result, moisture evaporation is difficult, and the drying efficiency is reduced [155].
Various atomization gases such as air, N2, Ar, and CO2 have been used for spray drying [156,157,158]. For liquid feed solutions containing flammable solvents, inert gases such as nitrogen are used as a drying medium to protect the products from degradation. Air is generally used as the drying gas in aqueous feed solutions.
DOE is essential in pharmaceutical spray drying for optimizing process parameters, ensuring product quality, reducing costs, supporting regulatory compliance, and enabling successful scale-up [159]. Implementing DOE in terms of factorial/response-surface designs shall overcome the limitations of the trial-and-error procedure while maintaining efficient use of limited resources and materials and revealing optimal process and formulation parameters. A summary of studies that utilized DoE in the spray drying process is presented in Table 2.

4. Mathematical and Computational Modeling in Spray Drying

4.1. Single Droplet Modeling

In the spray drying process, a single droplet was dried in two stages. In the first stage, droplet evaporation occurs, causing a decrease in the droplet diameter. In the second stage, a dry solid crust containing wet particles is produced and continuously dried until the required moisture content is achieved. Understanding the drying kinetics and mechanisms, such as evaporation, solid distribution, and skin formation, during droplet formation is essential for process control and microparticle characteristics.
The drying kinetics were modeled by solving the coupled heat and mass transfer equations. Mass coupling from the gas to the droplet occurs via evaporation, momentum transfer, and energy coupling via heat transfer. The heat and mass transfer between the droplets and the hot gas are expressed as follows:
m d C p d T d d t = h A d T g T d + d m d d t h f g
where md is the droplet mass, C p is the droplet heat capacity, T d is the droplet temperature, h is the heat transfer coefficient, A d is the droplet surface area, T g is the glass transition temperature, and h f g is the latent heat.
Specific equations can be used to calculate the heat transfer coefficient (hT) and mass transfer rate d m d d t . The heat transfer coefficient derived from the Ranz–Marshall equation can be written as:
N u = 2 + 0.6 R e d 1 2 P r 1 3
where Nu is the Nusselt number, k is the thermal conductivity of the gas, R e d the Reynolds number, and P r is the Prandtl number defined as:
P r = C p μ k t a
Equation (4) gives the mass transfer rate for evaporation between the droplet and gas.
d m d d t = k c A d Y S Y g
where d m d d t is the mass transfer rate, Y S is the saturation humidity, Y g is the gas humidity, and k c is the mass transfer coefficient obtained from the Sherwood number (Sh):
S h = k c d d D g = 2 + 0.6 R e d 1 2 P r 1 3
where S h is the Sherwood number, D g is the diffusion coefficient of water vapor in the gas phase, and S c is the Schmidt number (Sc).
S c = μ ρ g D g
where μ is the molecular viscosity of the fluid and ρ g is the density of the fluid.
During drying, a droplet may contain either insoluble or dissolved solids. There are some important factors, such as initial droplet heating in the first drying stage, temperature profile within the droplet, heat absorption by the crust region, crust resistance to diffusion mass transfer, and temperature dependence of the physical properties, that need to be considered for modeling the drying process of a droplet. A model was developed to simulate the drying of droplets containing suspended solids for the prediction of particle morphology, along with the droplet drying rate [179]. A population balance approach was used to model the discrete solid phase by assuming all particles are spherical. This approach predicts the critical solids volume fraction at which the shell forms, as illustrated in Figure 6. From the nature of the shell formed, particle morphology can be determined, as a shell formed from particles with a similar diameter will have a higher porosity than one consisting of particles of different sizes. In another study [180], an experimental technique using an acoustic levitator was carried out to examine the single droplet drying kinetics of spray-dried PLGA/ethyl acetate microparticles. The results presented the influence of process variables such as inlet temperature (0–40 °C), polymer concentration (5–400 mg/mL), and droplet size (0.5–1.37 μL) on the drying time and drying kinetics, as well as the particle morphology.
Previous studies mainly modeled the drying kinetics based on the average moisture content and neglected these factors [181,182]; however, a more advanced and realistic theoretical model that considered all these factors was developed, and it showed good agreement with experimental measurements, as shown in Figure 7A [183]. In a recent study, the single droplet drying kinetics of maltodextrin were investigated using experimental and numerical techniques. The experiments analyzed the variations in critical moisture content, and the model was employed for use in simulations of spray drying. However, the model accuracy was reported to be low [184]. Due to the low model accuracy of the CFD model, the optimized spray drying tower is expected to have uneven temperature distribution and particle trajectory, resulting in high wall deposition and low product quality. Therefore, a New CFD model (N-CFD) was developed to simulate the temperature distribution and particle trajectory in a spray drying chamber based on experimental data. Simulation results proposed a new type of air inlet mechanism that can even out the temperature distribution and particle trajectory in the drying chamber [185]. CFD simulations were also carried out in the food industry to analyze the effect of milk performance parameters on the spray dryer conditions. Results showed that increasing both the milk mass flow rate and the diameter of the particles resulted in an increase in particle velocity and a decrease in both the inlet temperature and outlet temperature of the particles [186].
Recently, a successful lumped drying model, the Reaction Engineering Approach (REA), was developed to dry individual droplets containing dissolved solids [187,188]. The REA model postulates that moisture must overcome an energy barrier to migrate from the interior to the exterior of the droplet surface. According to this approach, the partial vapor pressure over the surface of a droplet with dissolved solids is not only a function of temperature but also of the droplet’s moisture content [189]. The REA model was used to predict drying kinetics in terms of droplet size, shape, and morphological changes during the drying process. The developed model was then compared with experimental measurements. As depicted in Figure 7B, the droplet drying curves of both the REA model and experimental data are in good agreement.
Figure 7. Drying of a single droplet: (A) temporal evolution of moisture mass fraction (simulation versus experimental results) [179] with permission from Elsevier, 2025; (B) droplet drying curves of Polyethylene glycol 6000 in terms of dry mass content (X) versus time, theoretical prediction of the REA model versus experiments [188].
Figure 7. Drying of a single droplet: (A) temporal evolution of moisture mass fraction (simulation versus experimental results) [179] with permission from Elsevier, 2025; (B) droplet drying curves of Polyethylene glycol 6000 in terms of dry mass content (X) versus time, theoretical prediction of the REA model versus experiments [188].
Pharmaceutics 17 01605 g007

4.2. CFD in Spray Drying

CFD simulations are commonly used in spray drying to model and optimize the design parameters, thereby enhancing the product quality [190,191,192,193,194]. The complex nature of the spray drying process includes multiple physical phenomena, such as liquid atomization, droplet interactions, agglomeration and coalescence of the particles, and mass and heat transfer on the scale of single droplets. Using CFD, we can predict the gas flow patterns, temperature and velocity distributions, particle trajectories, and humidity within the drying chamber [190,195,196,197].
CFD analysis involves three stages: preprocessing, numerical simulation, and postprocessing. In the first stage, the geometry and boundary conditions are defined, and a mesh is generated for the drying chamber [198]. During the processing stage, partial differential equations are discretized using finite element, finite difference, or finite volume methods. Boundary conditions are applied, and the resulting algebraic equations for each mesh are numerically solved using Gaussian elimination or the Gauss–Seidel method are numerically solved. The residual errors between consecutive iterations are minimized until convergence is achieved. The final stage involves the visualization of the simulation results using different contour plots that allow us to make decisions for design optimization. All the stages involved in the CFD analysis are shown in Figure 8A. During the spray process, phenomena such as particle–particle collisions and particle–wall interactions can be observed inside the drying chamber, as depicted in Figure 9B. Accurate prediction of particle–particle collisions is important, as it might affect the size and morphology of the final dried powder. Conversely, particle–wall collisions result in a low powder yield owing to the deposition of particles on the walls of the drying chamber. The particle residence time distribution is crucial because of the heat-sensitive nature of some products. Table 3 presents an overview of the use of CFD modeling in a wide range of applications.
The spray drying process involves a gas phase (drying air), liquid phase (droplets), and solid phase (particles). Because the gas phase occupies more volume than the liquid and solid phases during spray drying, it can be assumed to be a continuous phase, whereas the solid and liquid phases can be considered discrete phases. The velocity and temperature profiles for the continuous phase transport can be predicted using CFD simulation, as illustrated in Figure 9A,B. For the continuous phase, the conservation of mass, momentum, energy, heat, and mass transfer, temperature, turbulence kinetic energy, and dissipation rate of kinetic energy are expressed as [208]:
Figure 9. CFD simulation of droplet drying in spray dryer: (A) 2D temperature profiles; (B) velocity profiles for continuous phase transport in a spray dryer; (C) droplet trajectory for average evaporation rate. Reproduced or adapted from [209], with permission from Elsevier, 2025.
Figure 9. CFD simulation of droplet drying in spray dryer: (A) 2D temperature profiles; (B) velocity profiles for continuous phase transport in a spray dryer; (C) droplet trajectory for average evaporation rate. Reproduced or adapted from [209], with permission from Elsevier, 2025.
Pharmaceutics 17 01605 g009
Continuity equation:
ρ t + x i ρ u i = 0
where ρ is the density, x is the cartesian coordinate, u is the velocity, and i is the cartesian coordinate index.
Momentum equation:
t ρ u i + x j ρ u i u j = x j p δ i j + μ u i x j + u j x i + ρ g i
where δ is the Kronecker delta, µ is dynamic viscosity, and g is the acceleration due to gravity.
Energy conservation:
t ρ C p T + x j ρ u j C p T x j K T x j = s T
where CP is the specific heat capacity, T is the temperature, K is the thermal conductivity, and sT is the thermal source.
The differential equation of heat transfer is used to determine the energy transfer and is given by:
t ρ C p u i = x j λ T x j + s T
Fick’s law of mass diffusion can be used to describe mass transfer and is expressed as follows:
X W t = x j D e f f X W x j
where Deff is the effective diffusion coefficient.
Turbulence kinetic energy:
t ρ k + x j ρ u j k = x j μ e σ k k x j + G k + G b ρ ε
Dissipation rate of turbulence kinetic energy:
t ρ ε + x j ρ u j ε = x j μ e σ ε ε x j + ε k C 1 G k C 2 ρ ε
For the discrete phase, particle/droplet movement is expressed by Newton’s Second Law:
d U p d t = g + F m p
Here, F is the sum of the forces (drag force (FD), buoyant force (FB), contact force (Fc), and added mass force (FA) acting on a given spray particle by the gas phase and other particles and walls of the spray drying chamber in Equation (14).
In the spray drying process, CFD mostly uses the Eulerian–Lagrangian method because it calculates the residence time of individual droplets within a wide range of droplet sizes [210]. In the CFD simulation, an Eulerian–Lagrangian model was used to calculate the droplet trajectories for different evaporation rates, as shown in Figure 9C by solving the force balance in Equation (15):
d u d d t = 18 μ ρ d d d 2 C D R e 24 v u d + g ρ d ρ g ρ d
where u d is the droplet velocity, ρ d is the droplet density, d d The droplet diameter, C D is the drag coefficient, Re is the Reynolds number, v is the fluid phase velocity, g is the gravitational force, and ρ g is the density of the fluid.
The droplet temperature is crucial for heat-sensitive products because it affects the thermal stability of heat-sensitive components. CFD simulation was used in the spray drying process to study the effect of droplet temperature in both short-form and tall-form spray dryers [210]. The droplet residence time significantly affects the critical quality attributes of the final powder, such as the solubility and bulk density. The droplet residence time in a superheated spray dryer was experimentally calculated and validated using CFD predictions. It has been reported that droplet residence time is significantly affected by droplet size and operating parameters [211]. Another study conducted in a short- and tall-form spray dryer suggested that most droplets had a low residence time during the spray drying process [210]. The droplet impact on the drying chamber is also significant for the design and operation of spray dryers and for improving the powder quality. The droplet impact positions were different for the short- and tall-form spray dryers. In the short-form spray dryer, a large proportion of the droplets struck the conical part of the drying chamber, whereas in the tall-form spray dryer, the droplets struck the cylindrical part [190].
Some of the CFD models for the spray drying process apply 3D analysis, an unsteady-state model, and a population balance approach to simulate the droplet drying mechanism inside the spray drying chamber to predict the asymmetry of the flow patterns inside the drying chamber [212]. Transient model calculations conducted for droplet–droplet interactions showed that droplet collisions affected temperature and humidity patterns, whereas their impact on velocity was less notable [213]. The population balance model explains droplet growth, coalescence, and breakup during spray drying [210]. It is frequently used to analyze dispersed systems such as colloids, polymers, and aerosols, and to track particle properties and their changes due to aggregation, fragmentation, nucleation, and growth in spray drying processes and various other fields [214].
The velocity profile, heat transfer, and air flow inside the spray drying chamber are extensively modeled using ANSYS Fluent 2025 R2 (commercial) and OpenFOAM v2506 (open source) platforms. These platforms provide high accuracy in turbulent modeling and have the capability to address domains with high spatial resolution, making them appropriate for detailed simulations in both academic and industrial environments. ANSYS Fluent is known for its user-friendly interface and robust physical models, making it accessible for users with little programming knowledge [215,216,217]. Nevertheless, it requires considerable resources and knowledge to function effectively. OpenFOAM is associated with a harder learning curve and generally requires more programming knowledge and skills to operate properly [216,217]. The key features of both platforms are summarized in Table 4.

4.3. Limitations and Recommendations in CFD Modeling

CFD modeling has several advantages, such as the prediction of temperature and velocity distribution, particle size, relative humidity, and several other factors of the spray drying process that influence the final properties of the dry powder [210]. However, there are also a few limitations owing to the coexistence of multiple phases (solids and fluids) in the spray drying process. In CFD spray drying simulations, model assumptions affect prediction accuracy. The choice of gas-phase turbulence model, steady-state or transient, and the selection of 2D axisymmetric vs. full 3D domains shall influence residence times, mixing, and recirculation patterns. Mesh discretization (including mesh quality and mesh size) and thermophysical simplifications (e.g., lumped thermal properties, neglected heat losses) further impact temperature and vapor distribution profiles. Reported maximum errors in discrete-phase predictions can approach 20% in such models [227]. Additional contributors include the treatment of particle–turbulence dispersion, atomizing air, initial atomization parameters (droplet size/velocity distributions), and wall heat-loss modeling [227,228]. It is difficult to predict the mass transfer that occurs within a particle during the spray–drying process using CFD modeling without the help of surrogate models that consider mass transfer, particle collisions, stickiness, and agglomeration. Moreover, experimental validation of CFD results is complicated and, to a certain extent, impossible owing to the limitations of experimental measurements [198,229], including a lack of adequate data on the physicochemical properties of materials, difficulty in determining the drying kinetics during the spray drying process, shrinkage during the drying process resulting in complex mesh generation, longer computational times owing to different time scales of fluid flow, heat, mass, and scalar transport, and different particle trajectories may arise owing to the circulation of drying air inside the drying chamber.
The drying kinetics and droplet shrinkage during the spray drying process are difficult to model using CFD modeling. A reduced-order model based on a population-based approach is an innovative method to accurately capture droplet shrinkage and drying kinetics. The main advantage of this approach is that it reduces computational costs, allowing it to simulate the entire production process. Furthermore, it uses the initial droplet size distribution to characterize the granules, thereby reducing the numerical errors due to discretization. The application of the population balance model in the spray-drying process is still limited. However, owing to its computational efficiency, it is recommended that the model accuracy be improved by including the droplet trajectory information from previously solved CFD simulations [212]. Particle size distribution is crucial in the spray-drying process. Droplet coalescence and particle agglomeration influence particle size distribution. Incorporating empirical expressions into an agglomeration/coalescence CFD model is an efficient way to accurately predict the particle size. To achieve this, existing binary collision models were modified and used in CFD together with a second adhesion model that involved more realistic droplet–particle, particle–droplet, and particle–particle collisions for efficient prediction of droplet–droplet collisions. Research studies [230,231] suggest that the proposed models can predict particle growth and final particle size with good accuracy and locally investigate the spray–drying process in different parts of the drying chamber. The main advantage of this CFD approach is that it considers heat loss and drying kinetics, along with particle agglomeration and coalescence. Studies related to droplet–droplet collisions have assumed only droplet coalescence for the prediction of particle agglomeration. Nevertheless, the prediction accuracy of the current model can be enhanced by further improving the sub-models for adhesion, formation of different droplet sizes, and particle breakage [230].
Accurate prediction of the residence time distribution (RTD) of droplets inside the spray drying chamber is crucial, as it influences the final powder properties. However, there is a lack of direct validation of experimental measurements with the CFD model, particularly in a counter-current spray dryer. Therefore, it is recommended to conduct specific studies that directly compare the measured and predicted RTD of droplets inside the drying chamber for enhancing the CQAs of the spray-dried powder [232]. Many CFD models assume that droplet shrinkage occurs because of water loss, thus neglecting the expansion/contraction of droplets due to trapped gases. These simplified assumptions can affect the prediction accuracy related to droplet drying kinetics and particle formation [233]. It is suggested to use more advanced drying kinetics models that consider the contraction and expansion of droplets for a more accurate representation of the particle formation process [228].
DEM combined with CFD was used to evaluate the role of surface tension on granule formation in a spray drying process. A CFD-DEM model containing a capillary force law based on Young–Laplace and Young equations was derived for particles in contact with liquid surfaces. Simulation results suggested that a non-linear correlation exists with subsequent granule density and morphology [234]. A coupled unresolved CFD-DEM approach was recently used for numerical simulation of the solidification of the single droplet suspension in a spray drying process. In this model, the Discrete Element Method (DEM) solves the equations of particle trajectory, while Computational Fluid Dynamics (CFD) computes the interstitial fluid flow [235]. The average drying conditions for a single droplet were obtained from a simulation of a large-scale spray dryer for different droplet sizes. The results showed that appropriate process parameters in a spray dryer can be determined using CFD-DEM simulation for the formation of particles with specific morphology [236].
A multi-scale multiphase approach was used to model the transport phenomena in a spray drying process based on experimental and theoretical data. The model utilized an Eulerian–Lagrangian approach and was solved numerically using the CFD technique. Successful validation and prediction of drying behavior of single droplets of suspensions, and two- and three-dimensional steady-state calculations of spray-drying processes were reported [237]. Another multi-scale approach that combines molecular-level packing interpretation with a continuum diffusion model was developed to predict the effect of initial droplet size on spray-dried protein-lactose powder. Model predictions pointed out that both the mean droplet size and the standard deviation of the log-normal size distribution had a huge influence on the surface composition of the particles [238].
Longer computational times owing to the different time spans of fluid flow, heat, mass, and scalar transport make the simulation process expensive. Previous studies [239] focused on reducing computational time by controlling the grid resolution by region. This approach focused only on the unsteady nature of the flow in different regions, and the grid resolution requirements changed dynamically within the domain. Therefore, using an advanced mesh refinement (AMR) strategy is an effective way to perform cost-effective transient simulations of the spray-drying process [240]. The use of AMR significantly reduced the computational cost compared to the base case (fixed grid with two refinement levels) and helped capture large eddies in critical regions. This approach provides a proper balance between solution accuracy and computational cost. Owing to the low computational cost involved, it is recommended to carry out parametric studies of the spray drying process, where computational expenses are a major limiting factor.
Several challenges in CFD simulations of spray drying can be addressed by implementing machine learning-aided models, hybrid models, and computation-efficient methods aided by machine learning. In the next section, the use of machine-learning-aided models to address the limitations of CFD models in the spray drying process is discussed in detail.

5. Machine Learning-Based Predictive Models

Machine learning (ML) techniques have recently emerged as integral tools for data analysis, predictive modeling, and decision-making in a wide range of engineering applications [241,242,243,244,245,246,247]. ML models can process large amounts of diverse data without vast computational resources [248]. To construct models that can generate accurate predictions, machine learning algorithms are trained on labeled datasets [249]. Important factors affecting model performance are the quality and reliability of the training data.
The core of the ML method comprises statistical and computational algorithms that analyze data, learn from it, and make predictions or decisions. The accuracy of the ML model is assessed quantitatively using performance metrics that determine the prediction level of the trained model [250].

5.1. Machine Learning Framework

The development of ML-based models consists of different stages, beginning with data preparation for model deployment, as depicted in Figure 10A. In the first stage, a dataset representing the studied problem–experimental or simulation trials–is prepared [251]. The collected dataset is then validated to evaluate the data quality for building an optimal ML model [252]. Second, different ML models are compared in terms of their prediction accuracy. In the third stage, hyperparameter tuning algorithms are applied to the trained model to obtain the optimal hyperparameters that generate the best prediction accuracy [253].
Subsequently, the model is evaluated by creating a confusion matrix to study the performance parameters such coefficient of determination (R2) and root mean squared error (RMSE). Finally, the best ML model was integrated into engineering applications. Regression algorithms were evaluated using metrics such as the coefficient of determination (R2), mean squared error (MSE), and root mean squared error (RMSE). A smaller RMSE value indicates better predictive accuracy:
Coefficient of Determination (R2):
R 2 = 1 i = 1 n y i y i ^ 2 i = 1 n y i y ¯ 2
MSE:
MSE = 1 n i = 1 n y i y i ^ 2
RMSE:
RMSE = 1 n i = 1 n y i y i ^ 2
where y i is the actual value, y i ^ the predicted value, y ¯ the mean of the actual values, and n the number of observations.
Various machine-learning algorithms are depicted in Figure 10B, ranging from basic linear regression to complex neural networks. The linear regression model is the simplest data-based ML model and is also referred to as the least-squares-based method [254], which is a direct method that predicts the response variable Y based on a single predictor variable X. It was assumed that a linear relationship exists between X and Y. If more than one predictor variable is present, the model is known as a multiple regression model.
Nonlinear regression models are frequently used for nonlinear relationships in complex systems [255]. Linear Support Vector Machines (SVMs) are supervised by ML algorithms that classify data by finding an optimal line or hyperplane that maximizes the distance between classes in an N-dimensional space [256,257]. They have been widely used in ML because they can handle both linear and nonlinear regressions and classifications [258,259]. SVMs allow for the generalization of new data and accurate classification and regression predictions.
Gaussian process models are generally used to carry out Bayesian nonlinear regression and classification and are integral to many machine-learning problems [260]. These models can model complex data relationships and identify high-level data properties, such as inputs, that are critical for identifying the response. Ensemble trees are another ML model that improves classification accuracy by combining multiple decision trees. It also addresses the overfitting and variance issues [261]. Variance refers to the sensitivity of a model to small changes in the training data. A high variance leads to overfitting of the target, whereas a low variance results in underfitting. Ensemble tree models include random forest, rotational forest, and extremely randomized trees [262]. Gradient Boosting Regression is another ensemble ML technique used for classification and regression. It builds a series of decision trees, and each new tree attempts to correct the errors made by previous ones. The final prediction was made by combining the outputs from all the trees [263].
Neural networks are ML-based empirical modeling techniques that can approximate continuous nonlinear relations [264]. These are the most complex ML models and contain artificial neurons that process and transmit signals [265]. They were trained using backpropagation and gradient descent methods, which contained strategies to eliminate overfitting. The different types of neural networks include feedforward, convolutional, and recurrent neural networks. Among neural networks, artificial neural networks (ANNs) are the most general type of neural network that can be used for different tasks such as classification, regression, and pattern recognition. ANNs have been studied in recent years and have shown better approximations than the RSM [266]. An artificial neural network is a computational method that develops data, such as data from the human brain [267]. It consists of an input layer that receives the initial data, an intermediate layer consisting of a complex structure of neurons (hidden layers), where complex mathematical calculations occur [268], and an output layer that produces results for the given input data. A schematic of the ANN is shown in Figure 11.
ML models have shown enormous potential in recognizing new CPPs that can effectively predict CQAs [100]. An overview of the use of ML models for predicting the CQAs of spray-dried powders is presented in Table 5.
The ML approach is used in various engineering applications and can be used to produce final biopharmaceutical products [269] and pharmaceutical spray drying [270,271,272,273]. In pharmaceutical drug delivery, ML plays a crucial role in improving drug solubility and bioavailability while minimizing experimental measurements. It utilizes huge experimental datasets and data–driven supervised algorithms for drug formulation optimization. In case of limited experimental data for model training, statistical learning models were employed to predict the properties and phenomena occurring between drug particles and carriers [274]. Multiple linear regression models were used to predict the long-term physical stability of poorly water-soluble drugs [275]. Recently, ML models such as ANNs have been prominently utilized in drug development for predicting and optimizing the drug composition, stability, and dissolution rates [276,277]. The particle size of the spray-dried powders is an important property affecting the bioavailability of the final drug product. An ensemble machine-learning model showed how the variation in process variables for a given pharmaceutical drug and formulation impacted the spray-dried particle size. In a recent study, ML models have been used for both experiments and image classification to examine the aerosol performance of dry powders during inhalation [269]. The powder yield is important for scale-up, and the median particle size distribution is crucial for processing and product design. Ensemble artificial neural networks can predict the powder yield and median particle size distribution of polymer-based amorphous solid dispersions [278]. Moreover, the ML approach has been successfully employed to optimize 3D-printed tablets [279], predict the fine-particle fraction and emitted dose of dry powder for inhalation [280], determine tablet defects [281], and determine the physical stability and dissolution rate of solid dispersions. The production of pharmaceutical drugs as amorphous formulations is considered a feasible method for improving the solubility of recent drug formulations [282]. However, amorphous solids tend to crystallize over time, which is difficult to predict. ML algorithms can be used to predict the physical stability of amorphous drugs [283]. Optimizing formulation parameters is crucial for ensuring the quality of liposomal nanoparticles [284]. Nevertheless, optimizing these parameters through experimentation is expensive and time-consuming [285]. ML models, through accurate predictions, reduce the cost and time associated with the spray drying process.
Table 5. ML models used to predict critical quality attributes of spray-dried powder.
Table 5. ML models used to predict critical quality attributes of spray-dried powder.
Machine Learning ModelsInput Parameters StudiedCritical Quality Attributes PredictedRefs.
Ensemble machine learning (EL)Sonication time, extrusion temperature, and feed compositionParticle size, and polydispersity index (PDI) of liposomal particles, in vitro dissolution profile[284,286]
Artificial neural networkDifferent types of drugs and excipients, carrier concentration, particle size, and morphologyDrug–excipient interactions, powder yield, emitted dose, fine particle fraction[280,286,287]
Support Vector Machine (SVM)Effects of the type of core and shell materials and their concentrations, effect of particle sizeIn vitro dissolution profile of sustained-release tablets, tablet tensile strength, and tablet brittleness index[286]
ML model accuracy can be improved by combining predictions from various learning models and provides a better assessment of optimum formulation parameters. A summary of the prediction accuracy of different ML models is given in Table 6.
Recently, ML has been suggested to assist in the Quality by Design process [291]. Further, experimental designs play a vital role in producing high-quality training datasets for ML models in the spray drying process. DoE offers systematic data collection and information related to the interaction between variables, whereas ML provides improved prediction capacity for process optimization and design space development. It has been proposed that predictive models developed based on DoE datasets using different ML algorithms, such as the Backpropagation Neural Network (BPNN), genetic algorithm-based BPNN, mind evolutionary algorithm-based BPNN, and Extreme gradient boosting. Machine have helped to understand the impact of critical process parameters that affect the final properties of the powder [292]. A combination of DoE and artificial neural networks as shown in Table 7 gave superior prediction performance compared to the traditional response surface methodology [293]. In another study, DoE was performed with a surrogate material such as α-lactose monohydrate instead of the actual protein. This was done to minimize the cost associated with the proteins. The prediction results of DoE and ML models were compared with three protein-based validation runs. Within the design space investigated, the powder yield, residual moisture content, and protein secondary structure were found to be satisfactory [294]. A thermodynamic model based on experimental points from the QbD approach was developed and compared with a CFD model to estimate the outlet temperature. The experimental points from the QbD approach were used to create the dataset for the thermodynamic model. The thermodynamic model utilized an ML approach to overcome the drawbacks of the lab-scale spray dryer. As a result, it showed a more accurate prediction of outlet temperature compared to the CFD model [295].
Data analysis, building, and deployment of ML models are carried out using ML platforms like TensorFlow 2.20.0 (open source) and MATLAB R2025a (commercial). TensorFlow is known for its robust and effective deep neural network (DNN) implementations. It is utilized to train advanced neural network models for prediction, control, and optimization tasks in spray drying process [296,297]. MATLAB is widely used for solving complex mathematical models and simulating drying kinetics, especially in 1–2D domain of the spray drying process. It has extensive built-in functions and is extremely useful for parameter optimization and data processing from CFD simulations [212,298]. The key elements of both platforms are discussed in Table 8.
The major challenge of industrial scalability of the spray drying process is designing and controlling the operating parameters, such as temperature, air flow, for achieving enhanced performance and quality of the final product. To overcome these limitations, modeling tools such as ML have emerged that can accurately predict complex relationships and nonlinear interactions between variables by training patterns. In the spray drying process, ML models provide crucial information about the transport phenomena through a porous material with certain features. Therefore, the implementation of AI in the industrial sector has made it possible to obtain products of the highest quality and reduced manufacturing costs throughout the entire product value chain [301]. In a milk spray drying process plant, the ML approach was utilized for the prediction and fault detection of important performance parameters. A neural network-based NARX model was employed in the prediction of cyclone exit air temperature, which is considered as a key performance parameter that affects thermal efficiency. The model accuracy for fault finding was reported as 99.83% [290].
ML technologies have been transforming various sectors, particularly the pharmaceutical industry. In pharmaceutical spray drying, ML can optimize batch production, allow predictive maintenance, enhance process control, and assist real-time quality monitoring [302]. Despite these benefits, regulatory agencies are aware of the unique threats of ML algorithms, such as continuous learning and uncertainty of the decision-making process, which pose a huge threat related to control, reproducibility, and traceability of the spray drying process [303]. This has tempted regulatory agencies to develop detailed frameworks and guidance (see Figure 12) for ML technologies that balance innovation, product quality, and patient adherence. Key regulatory challenges for implementing ML in the spray drying process include model validation and verification, data integrity, explainability and transparency, change management and lifecycle control, and ethical and legal considerations [304].
Model behavior may change over time. Therefore, regulatory authorities have advised a locked model during the time of validation, with a predetermined change control plan for any updates [305]. Data available must be reliable, accurate, and consistent for ML systems to utilize during training, testing, and deployment phases [306]. Moreover, the regulators demand that manufacturers understand the rational concept behind ML predictions and provide a clear justification based on scientific and engineering principles [307]. Further, regulatory authorities expect producers to define a model life cycle strategy that contains performance monitoring, retraining schedules, version control, and revalidation triggers. ML systems may accidentally introduce biases, particularly when trained on non-representative datasets. This may result in variable product quality during manufacturing. As a result, the regulatory authorities highlight the need for bias detection, fairness evaluation, and mitigation strategies as part of the model validation package [308].
The Food and Drug Administration (FDA) and European Medicines Agency (EMA) are the regulatory authorities that have been actively developing and releasing technical guidance and regulatory frameworks to address the incorporation of AI/ML in pharmaceutical drug delivery. A summary of key FDA/EMA elements is presented in Table 9.
Regulatory frameworks are not only about legal compliance, but they also promote trust, transparency, and fairness in AI systems. Explainability is considered crucial for promoting trust among stakeholders and clinicians, preventing algorithmic bias, and assuring that decisions are justifiable [314,315]. Therefore, explainable AI (XAI) is essential for transparency and explainability of ML models to the people, particularly in regulated domains such as the pharmaceutical industry. Regulatory authorities like the EU’s General Data Protection Regulation (GDPR) mandate that people affected by automated decisions have the right to get reasonable explanations about the ML models used in spray drying processes for drug manufacturing [316,317]. XAI helps users understand and trust model results by offering insights into the decision-making process, which is essential for regulatory authorities to evaluate the accuracy and integrity of ML-driven spray drying models [316,318]. The implementation of XAI promotes trust in AI systems by revealing how predictions are made, explaining relationships between input parameters and response variables, and quantitatively assessing the influence of the process parameters on predictions. This transparency is crucial for regulatory approval, as it facilitates ethical implementation and simplifies adherence to legal obligations [319,320]. In pharmaceutical drug manufacturing, explainable models can verify that decisions are made based on relevant trends, identify areas for improvement, and offer new perspectives on the data, all of which bolster regulatory review and approval [318,319].

5.2. Hybrid ML Models

The complexity of CFD simulation of the spray drying process increases with the level of physical details required, i.e., droplet atomization process, evaporation of single droplets, and modeling of the spray dryer unit. In the hybrid ML modeling approach, machine learning models are coupled with mechanistic-based model simulations. This integration allows benefiting from the advantages of both methods and eliminating their application shortcomings [321]. Hybrid ML models are commonly used to determine the optimal design of fundamental processes [322]. It also has applications in spray drying processes [323]. Hybrid ML models have been used for predicting the drug concentration in the spray drying process [324,325].
Combining ML and CFD models can be achieved via several approaches, as illustrated in Figure 13A CFD model is used to generate datasets for simulating the behavior of the physical phenomena under different experimental conditions (input parameters). This dataset is used by an ML algorithm to generate a trained ML model that can be used for prediction. In the second approach, as depicted in Figure 13B, the ML model is used as an input model (surrogate model) to the CFD model. The ML model is trained on data from physical experiments, simulations, or a combination of both. Therefore, it minimizes the need for repeated physical experiments. This approach was proposed to develop/optimize the spray drying process, thereby minimizing the requirement for costly experiments. Surrogate models are approximate mathematical models that closely mimic the behavior of a simulation model and are computationally efficient to evaluate. In a hybrid ML model, the data-driven model explains the system behavior based solely on data correlations [326], and the physical model provides a simplified representation of the real system that can be used to study, design, and test engineering problems. This hybrid approach helps in the development of an optimized algorithm by integrating the benefits of both physical and data-driven models. ML can help optimize parameters within the physical model, and the physical model can restrict the ML algorithm to prevent overfitting and ensure reasonable outcomes [327,328,329]. Therefore, by combining the two models, the hybrid ML model improves performance and reduces data requirements. It can also enhance prediction accuracy and decrease training complexity compared with single-algorithm approaches [330]. CFD combined with a mechanistic model was developed for a better understanding of the spray drying process of sticky materials. CFD model optimized the nozzle conditions by generating a large dataset of droplet size distribution for a subset of nozzles and spray angles. This data was directly fed into the mechanistic model for the prediction of heat and mass transfer between the gas and droplet, outlet temperature, size separation, residual solvent content in the particle, and droplet drying time [331]. The droplet drying kinetics of Lonicerae Japonicae Flos (LJF) extract were simulated using a CFD model to calculate the moisture removal rate of atomized droplets in the drying chamber. The CFD model was verified using experimental temperature field measurements. The experimental and simulation data were fed into a deep reinforcement machine learning algorithm for optimizing the spray drying process [332].
Energy efficiency is vital in the spray drying process as it can significantly improve the product yield, energy utilization, and minimize the operational costs. However, due to limited data issues, analyzing the performance of energy efficiency using an ML-based model is ineffective and inaccurate [334] To overcome these drawbacks, digital twin technology has recently emerged as a powerful tool for real-time process monitoring and decision-making by integrating with physical machines [335,336]. Digital twin approaches support the ML model by initiating parameters using limited simulated data from the source domain for effective prediction of energy efficiency, as shown in Figure 14. The digital twin-aided transfer learning model was verified in an industrial detergent powder spray drying process. Results suggest that the proposed model enhanced the drying efficiency by 14.53% and reduced the additional energy on demand and supply by 50.05% and 81.27%, respectively [337]. The digital twin approach delivers a comprehensive and real-time predictive view of process-equipment performance and prospective behavior, enabling early fault detection, remote troubleshooting, and improved operational reliability [338,339]. It can help in detecting possible defects, enabling remote troubleshooting, and eventually enhance customer satisfaction. Further, it helps with product distinction, product quality, and other supplementary services. The digital twin technique processes huge amounts of sensitive data, raising significant privacy concerns. Unauthorized access or data breaches can lead to operational interruptions and sacrifice critical infrastructure [340]. Lack of integrated management and decision-making strategies, along with partial monitoring and modeling, can leave digital twin systems exposed to cyber-attacks [341]. Adopting end-to-end encryption, regular security audits, and proactive threat detection mechanisms are crucial for protecting digital twin environments [342]. Real-time monitoring and strong asset authentication protocols are essential to maintain data integrity and respond quickly to potential breaches [339]. The effectiveness of digital twins depends largely on the quality and accuracy of the data they receive. Inconsistent data from sensors or other sources can sacrifice the reliability of the digital twin, leading to misleading insights and decision-making. Ensuring the authenticity of the data, particularly from the operational section, is essential for preserving the data integrity [341].

5.3. Limitations of Using ML/AI in Engineering

ML and AI have several advantages in engineering, including enhanced efficiency, improved decision-making, and advanced automation [241]. However, there are several limitations to combining ML and AI models in engineering systems. The key limitations include testing, AI software quality, and data management [343]. Other limitations include data quality issues, design method challenges, and performance concerns related to the ML model during the production stage [344]. AI and ML models largely depend on high-quality data to train the appropriate models; however, such data can be limited or difficult to obtain in complex system engineering [345]. Poor data quality can result in inaccurate predictions and decisions. Previous experimental data may also be limited, making it difficult to develop and train appropriate models [346]. Furthermore, the data may include historical biases, cultural stereotypes, or unequal representations of various groups, resulting in biased or discriminatory outcomes [347]. Biased algorithms can produce designs that do not effectively reflect the requirements of all users [348]. Combining AI/ML technologies with existing systems and software can be demanding, owing to compatibility issues resulting from differences in data formats, communication protocols, or architectural styles [346].

5.4. Comparative Analysis Between CFD and ML Models

CFD modeling provides detailed information about complex interactions between chemical and physical properties, fluid dynamics, and heat or mass transfer that occur during the spray drying process. It can simulate the droplet drying kinetics and their influence on powder properties, which is vital for pharmaceutical drug delivery [349]. Further, CFD models provide useful theoretical information for knowing the fluid flow patterns, droplet trajectory, and evaporation mechanism inside the spray drying tower. Advanced CFD-DEM models consider both fluid and particle interactions, which is significant for a spray drying process [350]. However, due to the complexity of the spray drying process, computational powder may be insufficient for industrial scalability. Most of the CFD models are focused on resolving fluid dynamics rather than on the final product output properties, also, experimental validation of CFD models is often limited [351].
ML models have been used to boost analytical models. In a recent study, the prediction accuracy of an ML model was compared with a simplified CFD model. The prediction accuracy of the ML model was six times better than that of the CFD model. Therefore, ML models can enhance the prediction accuracy, minimize the computational cost, and thereby speed up the process development. However, an ML model requires huge experimental datasets to make accurate predictions. Moreover, ML models are still an emerging technique, as most of the pharmaceutical spray drying applications have used ML models to assist the existing models and not fully replace them [352]. A comparative analysis of the CFD model and ML model in terms of different features is shown in Table 10.
A hybrid CFD-ML model combines the strengths of both the approaches and helps in making more accurate predictions than traditional CFD alone. However, ML models require reference data for training, emphasizing the need for advanced hybrid methods that enhance predictive power while reducing resource requirements Transfer learning is a well-established technique for mitigating data scarcity [353,354,355]. This approach adapts the knowledge gained from training a model on the source task to enhance performance on a different but related task (the target task). Instead of training a model from scratch for each new task, transfer learning adapts a pre-trained model to a new, often more specific, task. Further, this technique is particularly useful when the target task has limited data, as it involves reusing learned features, representations, or parameters from the source model to optimize the target model, leading to faster convergence and reduced computational cost and time.
Table 10. Summary Comparison between CFD and ML model.
Table 10. Summary Comparison between CFD and ML model.
CharacteristicsCFD ModelML ModelRefs.
Mechanistic insightSimulates physical phenomena, droplet drying kinetics, and process robustnessDepends on data patterns and has less physical insight[350,352]
Prediction accuracyGood: Based on model validation and computational resourcesHigher with substantial datasets[349,351,352]
Computational expenseHigh: Particularly for industrial scaleLow: predictions are rapid upon training[351,352]
Product quality predictionAble to model the influence of operating parameters on powder propertiesAble to enhance the prediction accuracy, but less mechanistic[349,352,356]
Industrial ApplicabilityLimited due to scale and validationPotential for automation and quick process optimization[351,352]

6. Conclusions

This review discusses the past and current state of research on modeling the spray drying process, particularly in the pharmaceutical industry. Spray drying is widely used in the food, chemical, and pharmaceutical industries. In the pharmaceutical industry, it is used for drug delivery via oral, nasal, transdermal, and inhalation routes. The spray-drying process involves heat and mass transfer, which results in the drying of droplets, followed by the formation of dry particles. Process parameters such as the inlet temperature, outlet temperature, drying gas flow rate, spray gas flow rate, feed concentration, and feed flow rate play a crucial role in determining the properties of the final dry powder. CFD modeling is commonly used in the spray-drying process for modeling the dryer unit. However, these models have many drawbacks, such as longer computational times, difficulty in experimental validation and determining drying kinetics, and inability to predict complex spray dryer conditions without the help of sub-models. ML models can be used to accurately predict spray drying process parameters and CQAs of the product. In this review study, we highlight recommendations concerning the challenges of CFD modeling in the spray drying process, which include the use of machine learning-based models, hybrid models, efficient computing methods, and validation practices. Coupling CFD models with novel modeling techniques, such as the reaction engineering approach and the drying kinetics model for enhancing model accuracy, using an advanced mesh refinement (AMR) strategy to perform cost-effective simulations, combining CFD models with DEM for determining the operating parameters, and utilizing a multi-scale approach to model the transport phenomena in a spray drying process. Advancements have been made in the application of ML models to the spray-drying process to overcome the limitations of CFD-based modelling. Digital twin-aided transfer learning is a recently emerged technique that assists ML models in making effective predictions using limited data. DoE datasets were also combined with ML models to improve prediction accuracy. Implementation of ML in the industrial sector is also possible through proper consideration of the regulatory framework. The FDA/EMA has released technical guidelines to address the implementation of AI/ML in the pharmaceutical sector. Explainable AI (XAI) is mandated by regulatory authorities to ensure the transparency and explainability of ML models to protect people affected by automated decisions. Nevertheless, ML models also have a few limitations, such as requiring high-quality datasets for better predictions, which are ineffective in predicting the influence of new process parameters on product quality. Moreover, this review outlines the lack of research utilizing the ML approach in spray drying, especially in the pharmaceutical field. Obtaining high-quality data consistently and using more advanced systems and software could improve the ML model prediction accuracy in the spray drying process. The utilization of the ML approach in the spray-drying process cannot be successful without proper experimental data to explain the CQAs of the final dried powder.

Author Contributions

W.W. conceptualized the study, provided critical revisions, and initial drafting of the manuscript. R.A. and B.A. contributed to data collection, literature review, and initial drafting of the manuscript. M.A. and Z.A. were involved in data analysis and figure preparation. B.A.-Z. acquired the funding and supervised the study and provided overall project guidance, contributed to the interpretation of results, and finalized the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the United Arab Emirates University, UAE, under research project grant No. G00004300 and G00004814. The authors therefore acknowledge with thanks the UAEU for technical and financial support.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AMRAdvanced Mesh Refinement
ANNArtificial Neural Network
APIActive pharmaceutical ingredient
CFDComputational Fluid Dynamics
CPPCritical process parameters
CQACritical quality attributes
DoEDesign of Experiment
DNNDeep neural network
DPIDry powder inhaler
EMAEuropean Medicines Agency
FDAFood and Drug Administration
GBRGradient Boosting Regression
GDPRGeneral Data Protection Regulation
MLMachine Learning
PePeclet number
QBDQuality by Design
REAReaction Engineering Approach
RMSERoot mean square error
RSMResponse Surface Methodology
RTDResidence time distribution
SVMSupport Vector Machine
XAIExplainable AI

References

  1. Sosnik, A.; Seremeta, K.P. Advantages and Challenges of the Spray-Drying Technology for the Production of Pure Drug Particles and Drug-Loaded Polymeric Carriers. Adv. Colloid Interface Sci. 2015, 223, 40–54. [Google Scholar] [CrossRef]
  2. Stefanescu, E.A.; Soto-Cantu, E. Modern Developments in the Spray-Drying Industries. Recent Pat. Mater. Sci. 2011, 4, 106–121. [Google Scholar] [CrossRef]
  3. Selvamuthukumaran, M. Handbook on Spray Drying Applications for Food Industries; CRC Press: Boca Raton, FL, USA, 2019; ISBN 9780429055133. [Google Scholar]
  4. Shishir, M.R.I.; Chen, W. Trends of Spray Drying: A Critical Review on Drying of Fruit and Vegetable Juices. Trends Food Sci. Technol. 2017, 65, 49–67. [Google Scholar] [CrossRef]
  5. Chauhan, A.K.; Patil, V. Effect of Packaging Material on Storage Ability of Mango Milk Powder and the Quality of Reconstituted Mango Milk Drink. Powder Technol. 2013, 239, 86–93. [Google Scholar] [CrossRef]
  6. Mohammed, N.K.; Tan, C.P.; Manap, Y.A.; Muhialdin, B.J.; Hussin, A.S.M. Spray Drying for the Encapsulation of Oils—A Review. Molecules 2020, 25, 3873. [Google Scholar] [CrossRef]
  7. Dantas, A.; Verruck, S.; Machado Canella, M.H.; Marchesan Maran, B.; Seigi Murakami, F.; de Avila Junior, B.; de Campos, C.E.M.; Hernandez, E.; Schwinden Prudencio, E. Current Knowledge about Physical Properties of Innovative Probiotic Spray-Dried Powders Produced with Lactose-Free Milk and Prebiotics. LWT 2021, 151, 112175. [Google Scholar] [CrossRef]
  8. Castro, F.; da Silva, N.R.; Silvério, S.C.; Ballesteros, L.F.; Teixeira, J.A. Chapter 17—Unit Operations for Extraction and Purification of Biological Products. In Current Developments in Biotechnology and Bioengineering; Sirohi, R., Pandey, A., Taherzadeh, M.J., Larroche, C., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 455–495. ISBN 978-0-323-91167-2. [Google Scholar]
  9. Erbay, Z.; Salum, P.; Bolat, E.B. 22—Drying of Dairy Products. In Drying Technology in Food Processing; Jafari, S.M., Malekjani, N., Eds.; Woodhead Publishing: London, UK, 2023; pp. 651–701. ISBN 978-0-12-819895-7. [Google Scholar]
  10. Oikonomopoulou, V.; Stramarkou, M.; Plakida, A.; Krokida, M. Optimization of Encapsulation of Stevia Glycosides through Electrospraying and Spray Drying. Food Hydrocoll. 2022, 131, 107854. [Google Scholar] [CrossRef]
  11. García-Carrasco, M.; Valdez-Baro, O.; Cabanillas-Bojórquez, L.A.; Bernal-Millán, M.J.; Rivera-Salas, M.M.; Gutiérrez-Grijalva, E.P.; Heredia, J.B. Potential Agricultural Uses of Micro/Nano Encapsulated Chitosan: A Review. Macromol 2023, 3, 614–635. [Google Scholar] [CrossRef]
  12. Merivaara, A.; Zini, J.; Koivunotko, E.; Valkonen, S.; Korhonen, O.; Fernandes, F.M.; Yliperttula, M. Preservation of Biomaterials and Cells by Freeze-Drying: Change of Paradigm. J. Control. Release 2021, 336, 480–498. [Google Scholar] [CrossRef]
  13. Rajashree, V.; Vijai Selvaraj, K.S.; Muthuvel, I.; Kavitha Shree, G.G.; Bharathi, A.; Senthil Kumar, M.; Rajiv, P. An Extensive Investigation of Combined Freeze-Drying Technologies for Fruit Conservation. Plant Sci. Today 2024, 11, 5515. [Google Scholar] [CrossRef]
  14. Ge, S.; Han, J.; Sun, Q.; Zhou, Q.; Ye, Z.; Li, P.; Gu, Q. Research Progress on Improving the Freeze-Drying Resistance of Probiotics: A Review. Trends Food Sci. Technol. 2024, 147, 104425. [Google Scholar] [CrossRef]
  15. Coşkun, N.; Sarıtaş, S.; Jaouhari, Y.; Bordiga, M.; Karav, S. The Impact of Freeze Drying on Bioactivity and Physical Properties of Food Products. Appl. Sci. 2024, 14, 9183. [Google Scholar] [CrossRef]
  16. Méndez-Lagunas, L.L.; Rodríguez-Ramírez, J.; Sandoval-Torres, S.; Barragán-Iglesias, J.; López-Ortíz, A. Strawberry Fluidized Bed Drying. Antiadhesion Pretreatments and Their Effect on Bioactive Compounds. Appl. Food Res. 2025, 5, 100739. [Google Scholar] [CrossRef]
  17. Setareh, M.; Assari, M.R.; Basirat Tabrizi, H.; Maghamian Zadeh, A. Experimental and Drying Kinetics Study on Millet Particles by a Pulsating Fluidized Bed Dryer. Heliyon 2024, 10, e33680. [Google Scholar] [CrossRef]
  18. Luthra, K.; Sadaka, S.S. Challenges and Opportunities Associated with Drying Rough Rice in Fluidized Bed Dryers: A Review. Trans. ASABE 2020, 63, 583–595. [Google Scholar] [CrossRef]
  19. Cheng, S.; Langrish, T.A.G. A Review of the Treatments to Reduce Anti-Nutritional Factors and Fluidized Bed Drying of Pulses. Foods 2025, 14, 681. [Google Scholar] [CrossRef]
  20. D’Addio, S.M.; Chan, J.G.Y.; Kwok, P.C.L.; Benson, B.R.; Prud’Homme, R.K.; Chan, H.K. Aerosol Delivery of Nanoparticles in Uniform Mannitol Carriers Formulated by Ultrasonic Spray Freeze Drying. Pharm. Res. 2013, 30, 2891–2901. [Google Scholar] [CrossRef]
  21. Wanning, S.; Süverkrüp, R.; Lamprecht, A. Jet-Vortex Spray Freeze Drying for the Production of Inhalable Lyophilisate Powders. Eur. J. Pharm. Sci. 2017, 96, 1–7. [Google Scholar] [CrossRef]
  22. Almakayeel, N.; Velu Kaliyannan, G.; Gunasekaran, R. Enhancing Power Conversion Efficiency of Polycrystalline Silicon Photovoltaic Cells Using Yttrium Oxide Anti-Reflective Coating via Electro-Spraying Method. Ceram. Int. 2024, 50, 42392–42403. [Google Scholar] [CrossRef]
  23. Hu, Z.; Hao, J.; Shen, D.; Gao, C.; Liu, Z.; Zhao, J.; Lu, B. Electro-Spraying/Spinning: A Novel Battery Manufacturing Technology. Green. Energy Environ. 2024, 9, 81–88. [Google Scholar] [CrossRef]
  24. Pandey, S.; Kumar, A.; Sharma, A. Sustainable Solar Drying: Recent Advances in Materials, Innovative Designs, Mathematical Modeling, and Energy Storage Solutions. Energy 2024, 308, 132725. [Google Scholar] [CrossRef]
  25. Tiwari, A. A Review on Solar Drying of Agricultural Produce. J. Food Process. Technol. 2016, 7, 1000623. [Google Scholar] [CrossRef]
  26. Bok, A. Advantages and Disadvantages of the Solar Drying of Sewage Sludge in Poland. Czas. Tech. 2017, 12, 171–179. [Google Scholar] [CrossRef][Green Version]
  27. Kherrafi, M.A.; Benseddik, A.; Saim, R.; Bouregueba, A.; Badji, A.; Nettari, C.; Hasrane, I. Advancements in Solar Drying Technologies: Design Variations, Hybrid Systems, Storage Materials and Numerical Analysis: A Review. Sol. Energy 2024, 270, 112383. [Google Scholar] [CrossRef]
  28. Jittanit, W.; Angkaew, K. Effect of Superheated-Steam Drying Compared to Conventional Parboiling on Chalkiness, Head Rice Yield and Quality of Chalky Rice Kernels. J. Stored Prod. Res. 2020, 87, 101627. [Google Scholar] [CrossRef]
  29. Speckhahn, A.; Srzednicki, G.; Desai, D.K. Drying of Beef in Superheated Steam. Dry. Technol. 2010, 28, 1072–1082. [Google Scholar] [CrossRef]
  30. Zhang, H.; Yu, F.; Yi, J.; Xu, X.; Ma, Y. Superheated Steam Technology: Recent Developments and Applications in Food Industries. Compr. Rev. Food Sci. Food Saf. 2024, 23, e70073. [Google Scholar] [CrossRef]
  31. Yan, Z.; Han, P.; Li, H.; Zheng, H.; Zhang, D. Superheated Steam Drying and Torrefaction of Wet Biomass: The Effect on Products Characteristics. Fuel 2025, 388, 134441. [Google Scholar] [CrossRef]
  32. Fu, H.; Yang, P.; Sun, P.; Ren, L.; Bu, Y.; Zhang, J.; Yang, D. Effects of Superheated Steam Processing Conditions on the Drying Behavior of Potatoes and Functional Properties of Potato Powder. Food Bioproc. Tech. 2025, 18, 9746–9762. [Google Scholar] [CrossRef]
  33. Obajemihi, O.I.; Cheng, J.-H.; Sun, D.-W. Novel Sequential and Simultaneous Infrared-Accelerated Drying Technologies for the Food Industry: Principles, Applications and Challenges. Crit. Rev. Food Sci. Nutr. 2023, 63, 1465–1482. [Google Scholar] [CrossRef]
  34. Akhoundzadeh Yamchi, A.; Hosainpour, A.; Hassanpour, A.; Rezvanivand Fanaei, A. Developing Ultrasound-Assisted Infrared Drying Technology for Bitter Melon (Momordica charantia). J. Food Process Eng. 2024, 47, e14516. [Google Scholar] [CrossRef]
  35. Song, J.; Jin, X.; Han, Y.; Zhai, S.; Zhang, K.; Jia, W.; Chen, J. Infrared Drying Effects on the Quality of Jujube and Process Optimization Using Response Surface Methodology. LWT 2024, 214, 117089. [Google Scholar] [CrossRef]
  36. Cansu, Ü. Utilization of Infrared Drying as Alternative to Spray- and Freeze-Drying for Low Energy Consumption in the Production of Powdered Gelatin. Gels 2024, 10, 522. [Google Scholar] [CrossRef]
  37. Pravallika, K.; Chakraborty, S.; Singhal, R.S. Supercritical Drying of Food Products: An Insightful Review. J. Food Eng. 2023, 343, 111375. [Google Scholar] [CrossRef]
  38. Zanotti, A.; Baldino, L.; Cardea, S.; Reverchon, E. Dye Removal from Wastewater Using Nanostructured Chitosan Aerogels Produced by Supercritical CO2 Drying. J. Supercrit. Fluids 2025, 216, 106442. [Google Scholar] [CrossRef]
  39. Demina, T.S.; Minaev, N.V.; Akopova, T.A. Polysaccharide-Based Aerogels Fabricated via Supercritical Fluid Drying: A Systematic Review. Polym. Bull. 2024, 81, 13331–13356. [Google Scholar] [CrossRef]
  40. Wu, H.T.; Li, T.H.; Tsai, H.M.; Chien, L.J.; Chuang, Y.H. Formulation of Inhalable Beclomethasone Dipropionate-Mannitol Composite Particles through Low-Temperature Supercritical Assisted Atomization. J. Supercrit. Fluids 2021, 168, 105095. [Google Scholar] [CrossRef]
  41. Machado, N.D.; Goñi, M.L.; Gañán, N.A. Effect of Supercritical CO2 Drying Variables and Gel Composition on the Textural Properties of Cellulose Aerogels. J. Supercrit. Fluids 2025, 215, 106414. [Google Scholar] [CrossRef]
  42. Rogers, T.L.; Johnston, K.P.; Williams, R.O., III. Solution-Based Particle Formation of Pharmaceutical Powders by Supercritical or Compressed Fluid CO2 and Cryogenic Spray-Freezing Technologies. Drug Dev. Ind. Pharm. 2001, 27, 1003–1015. [Google Scholar] [CrossRef]
  43. Snyder, H.E. Pharmaceutical Spray Drying: Solid-Dose Process Technology Platform for the 21st Century. Ther. Deliv. 2012, 3, 901–912. [Google Scholar] [CrossRef]
  44. Jain, M.S.; Lohare, G.B.; Bari, M.M.; Chavan, R.B.; Barhate, S.D.; Shah, C.B. Spray Drying in Pharmaceutical Industry: A Review. Res. J. Pharm. Dos. Forms Technol. 2012, 4, 74–79. [Google Scholar]
  45. Sollohub, K.; Cal, K. Spray Drying Technique: II. Current Applications in Pharmaceutical Technology. J. Pharm. Sci. 2010, 99, 587–597. [Google Scholar] [CrossRef]
  46. Patel, B.; Patel, J.; Chakraborty, S. Review of Patents and Application of Spray Drying in Pharmaceutical, Food and Flavor Industry. Recent. Pat. Drug Deliv. Formul. 2014, 8, 63–78. [Google Scholar] [CrossRef]
  47. Aundhia, C.J.; Raval, J.A.; Patel, M.M.; Shah, N.V.; Chauhan, S.P.; Sailor, G.U.; Javia, A.R. Spray Drying in the Pharmaceutical Industry—A Review. J. Pharm. Res. 2011, 2, 125–138. [Google Scholar]
  48. Lechanteur, A.; Evrard, B. Influence of Composition and Spray-Drying Process Parameters on Carrier-Free DPI Properties and Behaviors in the Lung: A Review. Pharmaceutics 2020, 12, 55. [Google Scholar] [CrossRef]
  49. Ré, M.-I. Formulating Drug Delivery Systems by Spray Drying. Dry. Technol. 2006, 24, 433–446. [Google Scholar] [CrossRef]
  50. Baumann, J.M.; Adam, M.S.; Wood, J.D. Engineering Advances in Spray Drying for Pharmaceuticals. Annu. Rev. Chem. Biomol. Eng. 2021, 12, 217–240. [Google Scholar] [CrossRef]
  51. Al Zaitone, B.; Al-Zahrani, A. Modeling Drying Behavior of an Aqueous Chitosan Single Droplet Using the Reaction Engineering Approach. AAPS PharmSciTech 2020, 21, 315. [Google Scholar] [CrossRef]
  52. Aliyar, H.; Schalau, G. Recent Developments in Silicones for Topical and Transdermal Drug Delivery. Ther. Deliv. 2015, 6, 827–839. [Google Scholar] [CrossRef]
  53. Bunker, M.J.; Roberts, C.J.; Davies, M.C.; James, M.B. A Nanoscale Study of Particle Friction in a Pharmaceutical System. Int. J. Pharm. 2006, 325, 163–171. [Google Scholar] [CrossRef]
  54. Gholizadeh-Hashjin, A.; Monajjemzadeh, F.; Ghafourian, T.; Hamishehkar, H.; Nokhodchi, A. Compatibility Study of Formoterol Fumarate-Lactose Dry Powder Inhalation Formulations: Spray Drying, Physical Mixture and Commercial DPIs. J. Drug Deliv. Sci. Technol. 2024, 95, 105538. [Google Scholar] [CrossRef]
  55. Varun, N.; Dutta, A.; Ghoroi, C. Influence of Surface Interaction between Drug and Excipient in Binary Mixture for Dry Powder Inhaler Applications. Adv. Powder Technol. 2022, 33, 103443. [Google Scholar] [CrossRef]
  56. de Boer, A.H.; Chan, H.K.; Price, R. A Critical View on Lactose-Based Drug Formulation and Device Studies for Dry Powder Inhalation: Which Are Relevant and What Interactions to Expect? Adv. Drug Deliv. Rev. 2012, 64, 257–274. [Google Scholar] [CrossRef] [PubMed]
  57. Kou, X.; Chan, L.W.; Steckel, H.; Heng, P.W.S. Physico-Chemical Aspects of Lactose for Inhalation. Adv. Drug Deliv. Rev. 2012, 64, 220–232. [Google Scholar] [CrossRef]
  58. Mehta, T.; Najafian, S.; Patel, K.; Lacombe, J.; Chaudhuri, B. Optimization of Carrier-Based Dry Powder Inhaler Performance: A Review. Pharmaceutics 2025, 17, 96. [Google Scholar] [CrossRef]
  59. Mahar, R.; Chakraborty, A.; Nainwal, N. The Influence of Carrier Type, Physical Characteristics, and Blending Techniques on the Performance of Dry Powder Inhalers. J. Drug Deliv. Sci. Technol. 2022, 76, 103759. [Google Scholar] [CrossRef]
  60. Mansour, H.M.; Xu, Z.; Hickey, A.J. Dry Powder Aerosols Generated by Standardized Entrainment Tubes from Alternative Sugar Blends: 3. Trehalose Dihydrate and d-Mannitol Carriers. J. Pharm. Sci. 2010, 99, 3430–3441. [Google Scholar] [CrossRef]
  61. Lechuga-Ballesteros, D.; Charan, C.; Stults, C.L.M.; Stevenson, C.L.; Miller, D.P.; Vehring, R.; Tep, V.; Kuo, M. Trileucine Improves Aerosol Performance and Stability of Spray-Dried Powders for Inhalation. J. Pharm. Sci. 2008, 97, 287–302. [Google Scholar] [CrossRef]
  62. Li, L.; Sun, S.; Parumasivam, T.; Denman, J.A.; Gengenbach, T.; Tang, P.; Mao, S.; Chan, H.-K. L-Leucine as an Excipient against Moisture on in Vitro Aerosolization Performances of Highly Hygroscopic Spray-Dried Powders. Eur. J. Pharm. Biopharm. 2016, 102, 132–141. [Google Scholar] [CrossRef]
  63. Littringer, E.M.; Mescher, A.; Schroettner, H.; Achelis, L.; Walzel, P.; Urbanetz, N.A. Spray Dried Mannitol Carrier Particles with Tailored Surface Properties—The Influence of Carrier Surface Roughness and Shape. Eur. J. Pharm. Biopharm. 2012, 82, 194–204. [Google Scholar] [CrossRef]
  64. Papakyriakopoulou, P.; Manta, K.; Kostantini, C.; Kikionis, S.; Banella, S.; Ioannou, E.; Christodoulou, E.; Rekkas, D.M.; Dallas, P.; Vertzoni, M.; et al. Nasal Powders of Quercetin-β-Cyclodextrin Derivatives Complexes with Mannitol/Lecithin Microparticles for Nose-to-Brain Delivery: In Vitro and Ex Vivo Evaluation. Int. J. Pharm. 2021, 607, 121016. [Google Scholar] [CrossRef] [PubMed]
  65. Duong, T.T.; Yen, T.T.H.; Nguyen, L.T.; Nguyen, T.D.; Nguyen, T.Q.T.; Nghiem, T.H.L.; Pham, H.T.; Raal, A.; Heinämäki, J.; Pham, T.M.H. Berberine-Loaded Liposomes for Oral Delivery: Preparation, Physicochemical Characterization and in-Vivo Evaluation in an Endogenous Hyperlipidemic Animal Model. Int. J. Pharm. 2022, 616, 121525. [Google Scholar] [CrossRef] [PubMed]
  66. Hu, J.; Dong, Y.; Ng, W.K.; Pastorin, G. Preparation of Drug Nanocrystals Embedded in Mannitol Microcrystals via Liquid Antisolvent Precipitation Followed by Immediate (on-Line) Spray Drying. Adv. Powder Technol. 2018, 29, 957–963. [Google Scholar] [CrossRef]
  67. Ma, Y.; Liu, Y.; Wang, Y.; Gao, P. Transdermal Codelivery System of Resveratrol Nanocrystals and Fluorouracil@ HP-β-CD by Dissolving Microneedles for Cutaneous Melanoma Treatment. J. Drug Deliv. Sci. Technol. 2024, 91, 105257. [Google Scholar] [CrossRef]
  68. Wang, B.; Friess, W. European Journal of Pharmaceutics and Biopharmaceutics Lipid-Coated Mannitol Core Microparticles for Sustained Release of Protein Placebo IgG Loaded Placebo. Eur. J. Pharm. Biopharm. 2018, 128, 91–97. [Google Scholar] [CrossRef]
  69. Mizoe, T.; Ozeki, T.; Okada, H. Preparation of Drug Nanoparticle-Containing Microparticles Using a 4-Fluid Nozzle Spray Drier for Oral, Pulmonary, and Injection Dosage Forms. J. Control. Release 2007, 122, 10–15. [Google Scholar] [CrossRef]
  70. Codrons, V.; Vanderbist, F.; Verbeeck, R.K.; Arras, M.; Lison, D.; Préat, V.; Vanbever, R. Systemic Delivery of Parathyroid Hormone (1-34) Using Inhalation Dry Powders in Rats. J. Pharm. Sci. 2003, 92, 938–950. [Google Scholar] [CrossRef]
  71. Ungaro, F.; D’Angelo, I.; Miro, A.; La Rotonda, M.I.; Quaglia, F. Engineered PLGA Nano- and Micro-Carriers for Pulmonary Delivery: Challenges and Promises. J. Pharm. Pharmacol. 2012, 64, 1217–1235. [Google Scholar] [CrossRef]
  72. Peng, T.; Lin, S.; Niu, B.; Wang, X.; Huang, Y.; Zhang, X.; Li, G.; Pan, X.; Wu, C. Influence of Physical Properties of Carrier on the Performance of Dry Powder Inhalers. Acta Pharm. Sin. B 2016, 6, 308–318. [Google Scholar] [CrossRef]
  73. Kaialy, W.; Ticehurst, M.; Nokhodchi, A. Dry Powder Inhalers: Mechanistic Evaluation of Lactose Formulations Containing Salbutamol Sulphate. Int. J. Pharm. 2012, 423, 184–194. [Google Scholar] [CrossRef]
  74. Trotta, V.; Lee, W.-H.; Loo, C.-Y.; Haghi, M.; Young, P.M.; Scalia, S.; Traini, D. In Vitro Biological Activity of Resveratrol Using a Novel Inhalable Resveratrol Spray-Dried Formulation. Int. J. Pharm. 2015, 491, 190–197. [Google Scholar] [CrossRef]
  75. Naikwade, S.R.; Bajaj, A.N.; Gurav, P.; Gatne, M.M.; Singh Soni, P. Development of Budesonide Microparticles Using Spray-Drying Technology for Pulmonary Administration: Design, Characterization, In Vitro Evaluation, and In Vivo Efficacy Study. AAPS PharmSciTech 2009, 10, 993–1012. [Google Scholar] [CrossRef] [PubMed]
  76. De Boeck, K.; Haarman, E.; Hull, J.; Lands, L.C.; Moeller, A.; Munck, A.; Riethmüller, J.; Tiddens, H.; Volpi, S.; Leadbetter, J.; et al. Inhaled Dry Powder Mannitol in Children with Cystic Fibrosis: A Randomised Efficacy and Safety Trial. J. Cyst. Fibros. 2017, 16, 380–387. [Google Scholar] [CrossRef] [PubMed]
  77. Bilton, D.; Robinson, P.; Cooper, P.; Gallagher, C.G.; Kolbe, J.; Fox, H.; Jaques, A.; Charlton, B. Inhaled Dry Powder Mannitol in Cystic Fibrosis: An Efficacy and Safety Study. Eur. Respir. J. 2011, 38, 1071–1080. [Google Scholar] [CrossRef] [PubMed]
  78. Duman, I.; Ünal, G.; Yilmaz, A.I.; Güney, A.Y.; Durduran, Y.; Pekcan, S. Inhaled Dry Powder Mannitol Treatment in Pediatric Patients with Cystic Fibrosis: Evaluation of Clinical Data in a Real-World Setting. Pediatr. Allergy Immunol. Pulmonol. 2022, 35, 19–26. [Google Scholar] [CrossRef]
  79. Marante, T.; Viegas, C.; Duarte, I.; Macedo, A.S.; Fonte, P. An Overview on Spray-Drying of Protein-Loaded Polymeric Nanoparticles for Dry Powder Inhalation. Pharmaceutics 2020, 12, 1032. [Google Scholar] [CrossRef]
  80. Anderson, S.D.; Daviskas, E.; Brannan, J.D.; Chan, H.K. Repurposing Excipients as Active Inhalation Agents: The Mannitol Story. Adv. Drug Deliv. Rev. 2018, 133, 45–56. [Google Scholar] [CrossRef]
  81. Mönckedieck, M.; Kamplade, J.; Fakner, P.; Urbanetz, N.A.; Walzel, P.; Steckel, H.; Scherließ, R. Dry Powder Inhaler Performance of Spray Dried Mannitol with Tailored Surface Morphologies as Carrier and Salbutamol Sulphate. Int. J. Pharm. 2017, 524, 351–363. [Google Scholar] [CrossRef]
  82. Torge, A.; Grützmacher, P.; Mücklich, F.; Schneider, M. The Influence of Mannitol on Morphology and Disintegration of Spray-Dried Nano-Embedded Microparticles. Eur. J. Pharm. Sci. 2017, 104, 171–179. [Google Scholar] [CrossRef]
  83. Zhang, W.; Gao, Y.; Wen, B.; Yang, T.; Liu, K.; Qin, M.; Ding, D. The Characteristics of Novel Quercetin Loaded Chitosan/Mannitol/Leucine Microspheres Prepared by Spray Drying for Inhalation. J. Control. Release 2017, 259, e137. [Google Scholar] [CrossRef]
  84. Har, C.L.; Fu, N.; Chan, E.S.; Tey, B.T.; Chen, X.D. In Situ Crystallization Kinetics and Behavior of Mannitol during Droplet Drying. Chem. Eng. J. 2018, 354, 314–326. [Google Scholar] [CrossRef]
  85. Almansour, K.; Alfagih, I.M.; Ali, R.; Elsayed, M.M.A. Inhalable Microparticles Containing Terbinafine for Management of Pulmonary Fungal Infections: Spray Drying Process Engineering Using Lactose vs. Mannitol as Excipients. J. Drug Deliv. Sci. Technol. 2020, 60, 101991. [Google Scholar] [CrossRef]
  86. Wu, J.; Wu, L.; Wan, F.; Rantanen, J.; Cun, D.; Yang, M. Effect of Thermal and Shear Stresses in the Spray Drying Process on the Stability of SiRNA Dry Powders. Int. J. Pharm. 2019, 566, 32–39. [Google Scholar] [CrossRef]
  87. Almansour, K.; Ali, R.; Alheibshy, F.; Almutairi, T.J.; Alshammari, R.F.; Alhajj, N.; Arpagaus, C.; Elsayed, M.M.A. Particle Engineering by Nano Spray Drying: Optimization of Process Parameters with Hydroethanolic versus Aqueous Solutions. Pharmaceutics 2022, 14, 800. [Google Scholar] [CrossRef]
  88. Ferdynand, M.S.; Nokhodchi, A. Co-Spraying of Carriers (Mannitol-Lactose) as a Method to Improve Aerosolization Performance of Salbutamol Sulfate Dry Powder Inhaler. Drug Deliv. Transl. Res. 2020, 10, 1418–1427. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, K.; Qian, S.; Liu, Z.; Liu, H.; Lin, Z.; Heng, W.; Gao, Y.; Zhang, J.; Wei, Y. Specific Surface Area of Mannitol Rather than Particle Size Dominant the Dissolution Rate of Poorly Water-Soluble Drug Tablets: A Study of Binary Mixture. Int J Pharm 2024, 660, 124280. [Google Scholar] [CrossRef] [PubMed]
  90. Yamasaki, K.; Kwok, P.C.L.; Fukushige, K.; Prud’Homme, R.K.; Chan, H.K. Enhanced Dissolution of Inhalable Cyclosporine Nano-Matrix Particles with Mannitol as Matrix Former. Int. J. Pharm. 2011, 420, 34–42. [Google Scholar] [CrossRef] [PubMed]
  91. Bosquillon, C.; Rouxhet, P.G.; Ahimou, F.; Simon, D.; Culot, C.; Préat, V.; Vanbever, R. Aerosolization Properties, Surface Composition and Physical State of Spray-Dried Protein Powders. J. Control. Release 2004, 99, 357–367. [Google Scholar] [CrossRef]
  92. Yu, D.; Hoag, S.W. The Impact of Diluents on the Compaction, Dissolution, and Physical Stability of Amorphous Solid Dispersion Tablets. Int. J. Pharm. 2024, 654, 123924. [Google Scholar] [CrossRef]
  93. Arte, K.S.; Tower, C.W.; Mutukuri, T.T.; Chen, Y.; Patel, S.M.; Munson, E.J.; Zhou, Q.T. Understanding the Impact of Mannitol on Physical Stability and Aerosolization of Spray-Dried Protein Powders for Inhalation. Int. J. Pharm. 2024, 650, 123698. [Google Scholar] [CrossRef]
  94. Ziaee, A.; Albadarin, A.B.; Padrela, L.; Femmer, T.; O’Reilly, E.; Walker, G. Spray Drying of Pharmaceuticals and Biopharmaceuticals: Critical Parameters and Experimental Process Optimization Approaches. Eur. J. Pharm. Sci. 2019, 127, 300–318. [Google Scholar] [CrossRef] [PubMed]
  95. Zbiciński, I.; Li, X. Conditions for Accurate CFD Modeling of Spray-Drying Process. Dry. Technol. 2006, 24, 1109–1114. [Google Scholar] [CrossRef]
  96. Bordón, M.G.; Alasino, N.P.X.; Camacho, N.M.; Millán-Rodríguez, F.; Pedroche-Jiménez, J.J.; Villanueva-Lazo, Á.; Ribotta, P.D.; Martínez, M.L. Mathematical Modeling of the Spray Drying Processes at Laboratory and Pilot Scales for the Development of Functional Microparticles Loaded with Chia Oil. Powder Technol. 2023, 430, 119018. [Google Scholar] [CrossRef]
  97. Focaroli, S.; Mah, P.T.; Hastedt, J.E.; Gitlin, I.; Oscarson, S.; Fahy, J.V.; Healy, A.M. A Design of Experiment (DoE) Approach to Optimise Spray Drying Process Conditions for the Production of Trehalose/Leucine Formulations with Application in Pulmonary Delivery. Int. J. Pharm. 2019, 562, 228–240. [Google Scholar] [CrossRef]
  98. Kanojia, G.; Willems, G.-J.; Frijlink, H.W.; Kersten, G.F.A.; Soema, P.C.; Amorij, J.-P. A Design of Experiment Approach to Predict Product and Process Parameters for a Spray Dried Influenza Vaccine. Int. J. Pharm. 2016, 511, 1098–1111. [Google Scholar] [CrossRef]
  99. Beg, S.; Swain, S.; Rahman, M.; Hasnain, M.S.; Imam, S.S. Chapter 3—Application of Design of Experiments (DoE). In Pharmaceutical Quality by Design: Principles and Applications; Academic Press: London, UK, 2019; pp. 43–64. ISBN 9780128157992. [Google Scholar]
  100. Walsh, I.; Myint, M.; Nguyen-khuong, T.; Ho, Y.S.; Kong, S.; Lakshmanan, M.; Walsh, I.; Myint, M.; Nguyen-khuong, T.; Ho, Y.S. Harnessing the Potential of Machine Learning for Advancing “ Quality by Design “ in Biomanufacturing. MAbs 2022, 14, 2013593. [Google Scholar] [CrossRef]
  101. Bazaria, B.; Kumar, P. Optimization of Spray Drying Parameters for Beetroot Juice Powder Using Response Surface Methodology (RSM). J. Saudi Soc. Agric. Sci. 2018, 17, 408–415. [Google Scholar] [CrossRef]
  102. Rahman, S.N.R.; Katari, O.; Pawde, D.M.; Boddeda, G.S.B.; Goswami, A.; Mutheneni, S.R.; Shunmugaperumal, T. Application of Design of Experiments® Approach-Driven Artificial Intelligence and Machine Learning for Systematic Optimization of Reverse Phase High Performance Liquid Chromatography Method to Analyze Simultaneously Two Drugs (Cyclosporin A and Etodolac) in Solution, Human Plasma, Nanocapsules, and Emulsions. AAPS PharmSciTech 2021, 22, 155. [Google Scholar] [CrossRef]
  103. Youssefi, S.; Emam-Djomeh, Z.; Mousavi, S.M. Comparison of Artificial Neural Network (ANN) and Response Surface Methodology (RSM) in the Prediction of Quality Parameters of Spray-Dried Pomegranate Juice. Dry. Technol. 2009, 27, 910–917. [Google Scholar] [CrossRef]
  104. Jariwala, N.; Putta, C.L.; Gatade, K.; Umarji, M.; Nazrin, S.; Rahman, R.; Pawde, D.M.; Sree, A.; Kamble, A.S.; Goswami, A.; et al. Journal of Drug Delivery Science and Technology Intriguing of Pharmaceutical Product Development Processes with the Help of Artificial Intelligence and Deep/Machine Learning or Artificial Neural Network. J. Drug Deliv. Sci. Technol. 2023, 87, 104751. [Google Scholar] [CrossRef]
  105. Allahyari, H.; Heidarpour, A.; Shayan, A.; Nguyen, V.P. A Robust Time-Dependent Model of Alkali-Silica Reaction at Different Temperatures. Cem. Concr. Compos. 2020, 106, 103460. [Google Scholar] [CrossRef]
  106. Zhang, S.; Gong, G.; Wang, H.; Chen, S.; Wang, J.; Qi, J. Advances and Challenges of the Spray Drying Technology: Towards Accurately Constructing Inorganic Multi-Functional Materials and Related Applications. Sustain. Mater. Technol. 2025, 43, e01250. [Google Scholar] [CrossRef]
  107. Bannigan, P.; Bao, Z.; Hickman, R.J.; Aldeghi, M.; Häse, F.; Aspuru-Guzik, A.; Allen, C. Machine Learning Models to Accelerate the Design of Polymeric Long-Acting Injectables. Nat. Commun. 2023, 14, 35. [Google Scholar] [CrossRef] [PubMed]
  108. Koleilat, L.; Feltner, L.; Stumpf, D.; Mort, P. Mapping Process-Product Relations by Ensemble Regression—A Granulation Case Study. Chem. Eng. Sci. 2025, 317, 122042. [Google Scholar] [CrossRef]
  109. Li, S.; Malvandi, A.; Feng, H.; Shao, C. Uncertainty-Aware Constrained Optimization for Air Convective Drying of Thin Apple Slices Using Machine-Learning-Based Response Surface Methodology. J. Food Eng. 2025, 394, 112503. [Google Scholar] [CrossRef]
  110. Walzel, P. Influence of the Spray Method on Product Quality and Morphology in Spray Drying. Chem. Eng. Technol. 2011, 34, 1039–1048. [Google Scholar] [CrossRef]
  111. Obón, J.M.; Luna-Abad, J.P.; Bermejo, B.; Fernández-López, J.A. Thermographic Studies of Cocurrent and Mixed Flow Spray Drying of Heat Sensitive Bioactive Compounds. J. Food Eng. 2020, 268, 109745. [Google Scholar] [CrossRef]
  112. Boel, E.; Koekoekx, R.; Dedroog, S.; Babkin, I.; Vetrano, M.R.; Clasen, C.; Van den Mooter, G. Unraveling Particle Formation: From Single Droplet Drying to Spray Drying and Electrospraying. Pharmaceutics 2020, 12, 625. [Google Scholar] [CrossRef]
  113. Dalla Barba, F.; Wang, J.; Picano, F. Revisiting D2-Law for the Evaporation of Dilute Droplets. Phys. Fluids 2021, 33, 51701. [Google Scholar] [CrossRef]
  114. Farid, M. A New Approach to Modelling of Single Droplet Drying. Chem. Eng. Sci. 2003, 58, 2985–2993. [Google Scholar] [CrossRef]
  115. Al Zaitone, B. Evaporation of Oblate Spheroidal Droplets: A Theoretical Analysis. Chem. Eng. Commun. 2018, 205, 110–121. [Google Scholar] [CrossRef]
  116. Ali Al Zaitone, B.; Tropea, C. Evaporation of Pure Liquid Droplets: Comparison of Droplet Evaporation in an Acoustic Field versus Glass-Filament. Chem. Eng. Sci. 2011, 66, 3914–3921. [Google Scholar] [CrossRef]
  117. Mezhericher, M.; Levy, A.; Borde, I. Theoretical Models of Single Droplet Drying Kinetics: A Review. Dry. Technol. 2010, 28, 278–293. [Google Scholar] [CrossRef]
  118. Dobry, D.E.; Settell, D.M.; Baumann, J.M.; Ray, R.J.; Graham, L.J.; Beyerinck, R.A. A Model-Based Methodology for Spray-Drying Process Development. J. Pharm. Innov. 2009, 4, 133–142. [Google Scholar] [CrossRef]
  119. Kim, E.H.J.; Chen, X.D.; Pearce, D. On the Mechanisms of Surface Formation and the Surface Compositions of Industrial Milk Powders. Dry. Technol. 2003, 21, 265–278. [Google Scholar] [CrossRef]
  120. Chaves, D.H.S.; Birchal, V.S.; Costa Junior, E.F. Particle Morphology and Shrinkage in Spray Dryers: A Review Focused on Modeling Single Droplet Drying. Dry. Technol. 2024, 42, 1138–1150. [Google Scholar] [CrossRef]
  121. Vasquez, L.; Buffo, A.; Vanni, M.; Frungieri, G. Micromechanics and Strength of Agglomerates Produced by Spray Drying. JCIS Open 2023, 9, 100068. [Google Scholar] [CrossRef]
  122. Alhajj, N.; Yahya, M.F.Z.R.; O’Reilly, N.J.; Cathcart, H. Development and Characterization of a Spray-Dried Inhalable Ternary Combination for the Treatment of Pseudomonas Aeruginosa Biofilm Infection in Cystic Fibrosis. Eur. J. Pharm. Sci. 2024, 192, 106654. [Google Scholar] [CrossRef] [PubMed]
  123. Sun, Y.; Cai, Z.; Fu, J. Particle Morphomics by High-Throughput Dynamic Image Analysis. Sci. Rep. 2019, 9, 9591. [Google Scholar] [CrossRef] [PubMed]
  124. Williams, D.S.; Pijpers, I.A.B.; Ridolfo, R.; van Hest, J.C.M. Controlling the Morphology of Copolymeric Vectors for next Generation Nanomedicine. J. Control. Release 2017, 259, 29–39. [Google Scholar] [CrossRef] [PubMed]
  125. Fang, H.; Cui, Z.; He, G.; Huang, L.; Chen, M. Phosphorus Adsorption onto Clay Minerals and Iron Oxide with Consideration of Heterogeneous Particle Morphology. Sci. Total Environ. 2017, 605–606, 357–367. [Google Scholar] [CrossRef]
  126. Al Zaitone, B.; Al-Zahrani, A. Spray Drying of Cellulose Nanofibers: Drying Kinetics Modeling of a Single Droplet and Particle Formation. Chem. Eng. Technol. 2021, 44, 1270–1277. [Google Scholar] [CrossRef]
  127. Vehring, R. Pharmaceutical Particle Engineering via Spray Drying. Pharm. Res. 2008, 25, 999–1022. [Google Scholar] [CrossRef] [PubMed]
  128. Edwards, D.A.; Dunbar, C.A. Bioengineering of Therapeutic Aerosols. Annu. Rev. Biomed. Eng. 2002, 4, 93–107. [Google Scholar] [CrossRef] [PubMed]
  129. Chow, A.H.L.; Tong, H.H.Y.; Chattopadhyay, P.; Shekunov, B. Particle Engineering for Pulmonary Drug Delivery. Pharm. Res. 2007, 24, 411–437. [Google Scholar] [CrossRef] [PubMed]
  130. Maa, Y.-F.; Henry R., C.; Phuong-Anh, N.; Hsu, C.C. The Effect of Operating and Formulation Variables on the Morphology of Spray-Dried Protein Particles. Pharm. Dev. Technol. 1997, 2, 213–223. [Google Scholar] [CrossRef]
  131. Schmitt, J.M.; Baumann, J.M.; Morgen, M.M. Predicting Spray Dried Dispersion Particle Size Via Machine Learning Regression Methods. Pharm. Res. 2022, 39, 3223–3239. [Google Scholar] [CrossRef]
  132. Krishna, K.V.M.; Anantatamukala, A.; Dahotre, N.B. Deep Learning Based Automated Quantification of Powders Used in Additive Manufacturing. Addit. Manuf. Lett. 2024, 11, 100241. [Google Scholar] [CrossRef]
  133. Muñoz López, C.A.; Peeters, K.; Van Impe, J. Data-Driven Modeling of the Spray Drying Process. Process Monitoring and Prediction of the Particle Size in Pharmaceutical Production. ACS Omega 2024, 9, 25678–25693. [Google Scholar] [CrossRef]
  134. Anis, S.; Urfi, S.M.; Kusumastuti, A.; Baskoro, W.W. Analysis of Inlet Temperature and Airflow Rate on Drying Process in a Spray Dryer Using Computational Fluid Dynamics Method. J. Adv. Res. Fluid Mech. Therm. Sci. 2022, 1, 163–171. [Google Scholar] [CrossRef]
  135. Langrish, T.; Chiou, D. Producing Powders of Hibiscus Extract in a Laboratory-Scale Spray Dryer. Int. J. Food Eng. 2008, 4. [Google Scholar] [CrossRef]
  136. Mohamad, G. Abiad, O.H.C.; Carvajal, M.T. Effect of Spray Drying Conditions on the Physicochemical Properties and Enthalpy Relaxation of α-Lactose. Int. J. Food Prop. 2014, 17, 1303–1316. [Google Scholar] [CrossRef]
  137. Littringer, E.M.; Noisternig, M.F.; Mescher, A.; Schroettner, H.; Walzel, P.; Griesser, U.J.; Urbanetz, N.A. The Morphology and Various Densities of Spray Dried Mannitol. Powder Technol. 2013, 246, 193–200. [Google Scholar] [CrossRef]
  138. Zier, K.-I.; Wulf, S.; Leopold, C.S. Factors Influencing the Properties and the Stability of Spray-Dried Sennae Fructus Extracts. Drug Dev. Ind. Pharm. 2018, 44, 1659–1667. [Google Scholar] [CrossRef]
  139. Both, E.M.; Boom, R.M.; Schutyser, M.A.I. Particle Morphology and Powder Properties during Spray Drying of Maltodextrin and Whey Protein Mixtures. Powder Technol. 2020, 363, 519–524. [Google Scholar] [CrossRef]
  140. Littringer, E.M.; Mescher, A.; Eckhard, S.; Schröttner, H.; Langes, C.; Fries, M.; Griesser, U.; Walzel, P.; Urbanetz, N.A. Spray Drying of Mannitol as a Drug Carrier—The Impact of Process Parameters on Product Properties. Dry. Technol. 2012, 30, 114–124. [Google Scholar] [CrossRef]
  141. Chiou, D.G.; Langrish, T.A.G.; Braham, R. The Effect of Temperature on the Crystallinity of Lactose Powders Produced by Spray Drying. J. Food Eng. 2008, 86, 288–293. [Google Scholar] [CrossRef]
  142. Tay, J.B.J.; Chua, X.; Ang, C.; Subramanian, G.S.; Tan, S.Y.; Lin, E.M.J.; Wu, W.Y.; Goh, K.K.T.; Lim, K. Effects of Spray-Drying Inlet Temperature on the Production of High-Quality Native Rice Starch. Processes 2021, 9, 1557. [Google Scholar] [CrossRef]
  143. Singh, A.; Van den Mooter, G. Spray Drying Formulation of Amorphous Solid Dispersions. Adv. Drug Deliv. Rev. 2016, 100, 27–50. [Google Scholar] [CrossRef]
  144. Zare, D.; Salehi, A.; Niakousari, M. Determination of Physical Properties of Sour Orange Juice Powder Produced by a Spray Dryer. In Proceedings of the American Society of Agricultural and Biological Engineers Annual International Meeting 2012, Dallas, TX, USA, 29 July–1 August 2012; Volume 5, pp. 4240–4253. [Google Scholar]
  145. Kathiman, M.N.; Abdul Mudalip, S.K.; Gimbun, J. Effect of Feed Flowrates on the Physical Properties and Antioxidant of Mahkota Dewa (Phaleria macrocarpa) Encapsulated Powder. IIUM Eng. J. 2022, 23, 1–9. [Google Scholar] [CrossRef]
  146. Maas, S.G.; Schaldach, G.; Littringer, E.M.; Mescher, A.; Griesser, U.J.; Braun, D.E.; Walzel, P.E.; Urbanetz, N.A. The Impact of Spray Drying Outlet Temperature on the Particle Morphology of Mannitol. Powder Technol. 2011, 213, 27–35. [Google Scholar] [CrossRef]
  147. Langrish, T.A.G. Assessing the Rate of Solid-Phase Crystallization for Lactose: The Effect of the Difference between Material and Glass-Transition Temperatures. Food Res. Int. 2008, 41, 630–636. [Google Scholar] [CrossRef]
  148. Anandharamakrishnan, C.; Rielly, C.D.; Stapley, A.G.F. Effects of Process Variables on the Denaturation of Whey Proteins during Spray Drying. Dry. Technol. 2007, 25, 799–807. [Google Scholar] [CrossRef]
  149. Zhang, Y.; Lin, J.; Zhong, Q. LWT—Food Science and Technology Effects of Media, Heat Adaptation, and Outlet Temperature on the Survival of Lactobacillus Salivarius NRRL B-30514 after Spray Drying and Subsequent Storage. LWT—Food Sci. Technol. 2016, 74, 441–447. [Google Scholar] [CrossRef]
  150. Littringer, E.M.; Paus, R.; Mescher, A.; Schroettner, H.; Walzel, P.; Urbanetz, N.A. The Morphology of Spray Dried Mannitol Particles—The Vital Importance of Droplet Size. Powder Technol. 2013, 239, 162–174. [Google Scholar] [CrossRef]
  151. Tonon, R.V.; Brabet, C.; Hubinger, M.D. Influence of Process Conditions on the Physicochemical Properties of Açai (Euterpe oleraceae Mart.) Powder Produced by Spray Drying. J. Food Eng. 2008, 88, 411–418. [Google Scholar] [CrossRef]
  152. Vickovic, D.; Czaja, T.P.; Gaiani, C.; Pedersen, S.J.; Ahrné, L.; Hougaard, A.B. The Effect of Feed Formulation on Surface Composition of Powders and Wall Deposition during Spray Drying of Acidified Dairy Products. Powder Technol. 2023, 418, 118297. [Google Scholar] [CrossRef]
  153. Goula, A.M.; Adamopoulos, K.G. Spray Drying of Tomato Pulp: Effect of Feed Concentration. Dry. Technol. 2004, 22, 2309–2330. [Google Scholar] [CrossRef]
  154. de Oliveira, A.H.; Mata, M.E.R.M.C.; Fortes, M.; Duarte, M.E.M.; Pasquali, M.; Lisboa, H.M. Influence of Spray Drying Conditions on the Properties of Whole Goat Milk. Dry. Technol. 2020, 39, 726–737. [Google Scholar] [CrossRef]
  155. Tan, L.W.; Ibrahim, M.N.; Kamil, R.; Taip, F.S. Empirical Modeling for Spray Drying Process of Sticky and Non-Sticky Products. Procedia Food Sci. 2011, 1, 690–697. [Google Scholar] [CrossRef]
  156. Szczap, J.P.; Jacobs, I.C. Atomization and Spray Drying Processes. In Microencapsulation in the Food Industry: A Practical Implementation Guide; Academic Press: London, UK, 2023; pp. 59–71. ISBN 9780128216835. [Google Scholar]
  157. Moura, C.; Casimiro, T.; Costa, E.; Aguiar-Ricardo, A. Optimization of Supercritical CO2-Assisted Spray Drying Technology for the Production of Inhalable Composite Particles Using Quality-by-Design Principles. Powder Technol. 2019, 357, 387–397. [Google Scholar] [CrossRef]
  158. Islam, M.I.U.; Langrish, T.A.G. The Effect of Different Atomizing Gases and Drying Media on the Crystallization Behavior of Spray-Dried Powders. Dry. Technol. 2010, 28, 1035–1043. [Google Scholar] [CrossRef]
  159. Spicher, M.T.; Schwaminger, S.P.; von der Haar-Leistl, D.; Fellenberg, E.; Berensmeier, S. Process Development for Pilot-Scale Spray Drying of Ultrasmall Iron (Oxyhydr)Oxide Nanoparticles. Powder Technol. 2024, 433, 119186. [Google Scholar] [CrossRef]
  160. LeClair, D.A.; Cranston, E.D.; Xing, Z.; Thompson, M.R. Optimization of Spray Drying Conditions for Yield, Particle Size and Biological Activity of Thermally Stable Viral Vectors. Pharm. Res. 2016, 33, 2763–2776. [Google Scholar] [CrossRef]
  161. Gil-Chávez, J.; Padhi, S.S.P.; Hartge, U.; Heinrich, S.; Smirnova, I. Optimization of the Spray-Drying Process for Developing Aquasolv Lignin Particles Using Response Surface Methodology. Adv. Powder Technol. 2020, 31, 2348–2356. [Google Scholar] [CrossRef]
  162. Ayre, A.; Lalitha, K.G.; Ruckmani, K.; Khutle, N.; Pawar, H.; Dand, N.; Vijaya, C. ICH Q8 Guidelines in Practice: Spray Drying Process Optimization by 23 Factorial Design for the Production of Famotidine Nanoparticles. Pharm. Nanotechnol. 2014, 2, 138–148. [Google Scholar] [CrossRef]
  163. Thakkar, V.; Pandey, E.; Pandya, T.; Shah, P.; Patel, A.; Trivedi, R.; Gohel, M.; Baldaniya, L.; Gandhi, T. Formulation of Dry Powder Inhaler of Anti-Tuberculous Drugs Using Spray Drying Technique and Optimization Using 23 Level Factorial Design Approach. Curr. Drug Ther. 2019, 14, 239–260. [Google Scholar] [CrossRef]
  164. Saha, T.; Sinha, S.; Harfoot, R.; Quiñones-Mateu, M.E.; Das, S.C. Manipulation of Spray-Drying Conditions to Develop an Inhalable Ivermectin Dry Powder. Pharmaceutics 2022, 14, 1432. [Google Scholar] [CrossRef]
  165. Narayan, P.; Marchant, D.; Wheatley, M.A. Optimization of Spray Drying by Factorial Design for Production of Hollow Microspheres for Ultrasound Imaging. J. Biomed. Mater. Res. 2001, 56, 333–341. [Google Scholar] [CrossRef] [PubMed]
  166. Guimarães, T.F.; Lanchote, A.D.; Da Costa, J.S.; Viçosa, A.L.; De Freitas, L.A.P. A Multivariate Approach Applied to Quality on Particle Engineering of Spray-Dried Mannitol. Adv. Powder Technol. 2015, 26, 1094–1101. [Google Scholar] [CrossRef]
  167. Benjasirimongkol, P.; Piriyaprasarth, S.; Sriamornsak, P. Effect of Formulations and Spray Drying Process Conditions on Physical Properties of Resveratrol Spray-Dried Emulsions. Key Eng. Mater. 2019, 819, 246–251. [Google Scholar] [CrossRef]
  168. Rathod, V.; Gajera, B.; Pinninti, A.; Mohammed, I.A.; Dave, R.H. Strategizing Spray Drying Process Optimization for the Manufacture of Redispersible Indomethacin Nanoparticles Using Quality-by-Design Principles. AAPS PharmSciTech 2023, 24, 133. [Google Scholar] [CrossRef]
  169. Karas, J.; Pavloková, S.; Hořavová, H.; Gajdziok, J. Optimization of Spray Drying Process Parameters for the Preparation of Inhalable Mannitol-Based Microparticles Using a Box-Behnken Experimental Design. Pharmaceutics 2023, 15, 496. [Google Scholar] [CrossRef] [PubMed]
  170. Kramek-Romanowska, K.; Odziomek, M.; Sosnowski, T.R.; Gradoń, L. Effects of Process Variables on the Properties of Spray-Dried Mannitol and Mannitol/Disodium Cromoglycate Powders Suitable for Drug Delivery by Inhalation. Ind. Eng. Chem. Res. 2011, 50, 13922–13931. [Google Scholar] [CrossRef]
  171. Amaro, M.I.; Tajber, L.; Corrigan, O.I.; Healy, A.M. Optimisation of Spray Drying Process Conditions for Sugar Nanoporous Microparticles (NPMPs) Intended for Inhalation. Int. J. Pharm. 2011, 421, 99–109. [Google Scholar] [CrossRef]
  172. Anish, C.; Upadhyay, A.K.; Sehgal, D.; Panda, A.K. Influences of Process and Formulation Parameters on Powder Flow Properties and Immunogenicity of Spray Dried Polymer Particles Entrapping Recombinant Pneumococcal Surface Protein A. Int. J. Pharm. 2014, 466, 198–210. [Google Scholar] [CrossRef]
  173. Saboo, S.; Tumban, E.; Peabody, J.; Wafula, D.; Peabody, D.S.; Chackerian, B.; Muttil, P. Optimized Formulation of a Thermostable Spray-Dried Virus-Like Particle Vaccine against Human Papillomavirus. Mol. Pharm. 2016, 13, 1646–1655. [Google Scholar] [CrossRef]
  174. Xia, D.; Shrestha, N.; van de Streek, J.; Mu, H.; Yang, M. Spray Drying of Fenofibrate Loaded Nanostructured Lipid Carriers. Asian J. Pharm. Sci. 2016, 11, 507–515. [Google Scholar] [CrossRef]
  175. Kouteu, P.A.N.; Kemegne, A.G.; Pahane, M.M.; Fanta, A.S.Y.; Membangmi, A.K.; Tchamaleu, D.V.T.; Agbor, G.A.; Mouangue, M.R. Face-Centered Central Composite Design for the Optimization of the Extraction of Phenolic Compounds from Kernels and Shells of Raphia Farinifera and Evaluation of the Antioxidant, Antimicrobial, and Anti-Inflammatory Activities. J. Food Process Preserv. 2024, 2024, 8849005. [Google Scholar] [CrossRef]
  176. Kauppinen, A.; Broekhuis, J.; Grasmeijer, N.; Tonnis, W.; Ketolainen, J.; Frijlink, H.W.; Hinrichs, W.L.J. Efficient Production of Solid Dispersions by Spray Drying Solutions of High Solid Content Using a 3-Fluid Nozzle. Eur. J. Pharm. Biopharm. 2018, 123, 50–58. [Google Scholar] [CrossRef]
  177. Ingvarsson, P.T.; Schmidt, S.T.; Christensen, D.; Larsen, N.B.; Hinrichs, W.L.J.; Andersen, P.; Rantanen, J.; Nielsen, H.M.; Yang, M.; Foged, C. Designing CAF-Adjuvanted Dry Powder Vaccines: Spray Drying Preserves the Adjuvant Activity of CAF01. J. Control. Release 2013, 167, 256–264. [Google Scholar] [CrossRef]
  178. Jensen, D.M.K.; Cun, D.; Maltesen, M.J.; Frokjaer, S.; Nielsen, H.M.; Foged, C. Spray Drying of SiRNA-Containing PLGA Nanoparticles Intended for Inhalation. J. Control. Release 2010, 142, 138–145. [Google Scholar] [CrossRef]
  179. Handscomb, C.S.; Kraft, M.; Bayly, A.E. A New Model for the Drying of Droplets Containing Suspended Solids. Chem. Eng. Sci. 2009, 64, 628–637. [Google Scholar] [CrossRef]
  180. Al Zaitone, B.; Lamprecht, A. Single Droplet Drying Step Characterization in Microsphere Preparation. Colloids Surf. B Biointerfaces 2013, 105, 328–334. [Google Scholar] [CrossRef]
  181. Kuts, P.S.; Strumillo, C.; Zbicinski, I. Evaporation Kinetics of Single Droplets Containing Dissolved Bioaiass. Dry. Technol. 1996, 14, 2041–2060. [Google Scholar] [CrossRef]
  182. Mezhericher, M.; Levy, A.; Borde, I. Theoretical Drying Model of Single Droplets Containing Insoluble or Dissolved Solids. Dry. Technol. 2007, 25, 1025–1032. [Google Scholar] [CrossRef]
  183. Griffith, J.D.; Bayly, A.E.; Johns, M.L. Magnetic Resonance Studies of Detergent Drop Drying. Chem. Eng. Sci. 2008, 63, 3449–3456. [Google Scholar] [CrossRef]
  184. Bürger, J.V.; Jaskulski, M.; Kharaghani, A. Modeling of Maltodextrin Drying Kinetics for Use in Simulations of Spray Drying. Dry. Technol. 2025, 43, 214–227. [Google Scholar] [CrossRef]
  185. Xiong, F.; Yuan, Y.; Liu, K.; Tan, L.; Li, J.; Zhao, Z. CFD Modeling of Baijiu Yeast Spray Drying Process and Improved Design of Drying Tower. Dry. Technol. 2024, 42, 836–853. [Google Scholar] [CrossRef]
  186. Ghafori, H.; Karimian, H. CFD Simulation of Milk Performance Parameters Effect on the Physical Indices of a Spray Dryer. Innov. Food Technol. 2025, 12, 163–181. [Google Scholar] [CrossRef]
  187. Chen, X.D.; Putranto, A. Reaction Engineering Approach (REA) to Modeling Drying Problems: Recent Development and Implementations. Dry. Technol. 2015, 33, 1899–1910. [Google Scholar] [CrossRef]
  188. Al Zaitone, B.; Al-Zahrani, A.; Ahmed, O.; Saeed, U.; Taimoor, A.A. Spray Drying of PEG6000 Suspension: Reaction Engineering Approach (REA) Modeling of Single Droplet Drying Kinetics. Processes 2022, 10, 1365. [Google Scholar] [CrossRef]
  189. Chen, X.D.; Lin, S.X.Q. Air Drying of Milk Droplet under Constant and Time-dependent Conditions. Aiche J. 2005, 51, 1790–1799. [Google Scholar] [CrossRef]
  190. Kuriakose, R.; Anandharamakrishnan, C. Computational Fluid Dynamics (CFD) Applications in Spray Drying of Food Products. Trends Food Sci. Technol. 2010, 21, 383–398. [Google Scholar] [CrossRef]
  191. Gimbun, J.; Law, W.P.; Anandharamakrishnan, C. Computational Fluid Dynamics Modelling of the Dairy Drying Processes. In Handbook of Drying for Dairy Products; Wiley-Blackwell: Hoboken, NJ, USA, 2017; pp. 179–201. ISBN 9781118930526. [Google Scholar]
  192. Huang, L.; Mujumdar, A.S. Numerical Study of Two-Stage Horizontal Spray Dryers Using Computational Fluid Dynamics. Dry. Technol. 2006, 24, 727–733. [Google Scholar] [CrossRef]
  193. Jamaleddine, T.J.; Ray, M.B. Application of Computational Fluid Dynamics for Simulation of Drying Processes: A Review. Dry. Technol. 2010, 28, 120–154. [Google Scholar] [CrossRef]
  194. Birchal, V.S.; Huang, L.; Mujumdar, A.S.; Passos, M.L. Spray Dryers: Modeling and Simulation. Dry. Technol. 2006, 24, 359–371. [Google Scholar] [CrossRef]
  195. Huang, L.; Kurichi, K.; Mujumdar, A.S. A Parametric Study of the Gas Flow Patterns and Drying Performance of Co-Current Spray Dryer: Results of a Computational Fluid Dynamics Study. Dry. Technol. 2003, 21, 957–978. [Google Scholar] [CrossRef]
  196. Rindang, A.; Panggabean, S.; Wulandari, F. CFD Analysis of Temperature Drying Chamber at Rotary Dryer With Combined Energy. J. Phys. Conf. Ser. 2019, 1155, 12037. [Google Scholar] [CrossRef]
  197. Huang, L.; Kurichi, K.; Mujumdar, A.S. Simulation of a Spray Dryer Fitted with a Rotary Disk Atomizer Using a Three-Dimensional Computional Fluid Dynamic Model. Dry. Technol. 2004, 22, 1489–1515. [Google Scholar] [CrossRef]
  198. Malekjani, N.; Jafari, S.M. Simulation of Food Drying Processes by Computational Fluid Dynamics (CFD); Recent Advances and Approaches. Trends Food Sci. Technol. 2018, 78, 206–223. [Google Scholar] [CrossRef]
  199. Salmanipour, S.; Sokhansanj, A.; Jafari, N.; Hamishehkar, H.; Saha, S.C. Engineering Nanoliposomal Tiotropium Bromide Embedded in a Lactose-Arginine Carrier Forming Trojan-Particle Dry Powders for Efficient Pulmonary Drug Delivery: A Combined Approach of in Vitro-3D Printing and in Silico-CFD Modeling. Int. J. Pharm. 2025, 671, 125171. [Google Scholar] [CrossRef] [PubMed]
  200. Coimbra, J.C.; Lopes, L.C.; da Silva Cotrim, W.; Prata, D.M. CFD Modeling of Spray Drying of Fresh Whey: Influence of Inlet Air Temperature on Drying, Fluid Dynamics, and Performance Indicators. Digit. Chem. Eng. 2024, 12, 100178. [Google Scholar] [CrossRef]
  201. Qomariyah, L.; Rahmana Putra, N.; Shamsuddin, R.; Hirano, T.; Ratna Puri, N.; Alawiyah, L.; Hamzah, A.; Sabar, S. Insights into Nanostructured Silica Particle Formation from Sodium Silicate Using Spray Drying: An Experimental and Theoretical Study. Chem. Phys. 2024, 585, 112347. [Google Scholar] [CrossRef]
  202. Zhang, F.; Lu, Y.; Zhu, S.; Han, F.; Wen, G.; Tang, P.; Hou, Z. Mathematical Model and Numerical Investigation of the Influence of Spray Drying Parameters on Granule Sizes of Mold Powder. Particuology 2024, 85, 280–295. [Google Scholar] [CrossRef]
  203. Aldo Benavides-Morán, A.C.; Gómez, A. Spray Drying Experiments and CFD Simulation of Guava Juice Formulation. Dry. Technol. 2021, 39, 450–465. [Google Scholar] [CrossRef]
  204. Qomariyah, L.; Puri, N.R.; Grady, E.; Nurtono, T.; Widiyastuti; Kusdianto; Madhania, S.; Winardi, S. Computational Fluid Dynamics (CFD) Modelling of ZnO-SiO2 Composite Through a Consecutive Electrospray and Spray Drying Method. J. Phys. Conf. Ser. 2022, 2344, 12004. [Google Scholar] [CrossRef]
  205. Lerotholi, L.; Everson, R.C.; Hattingh, B.B.; Koech, L.; Le Roux, I.; Neomagus, H.W.J.P.; Rutto, H. Computational Fluid Dynamics Modeling and Analysis of Lime Slurry Drying in a Laboratory Spray Dry Scrubber. Ind. Eng. Chem. Res. 2024, 63, 21038–21061. [Google Scholar] [CrossRef]
  206. Longest, P.W.; Farkas, D.; Hassan, A.; Hindle, M. Computational Fluid Dynamics (CFD) Simulations of Spray Drying: Linking Drying Parameters with Experimental Aerosolization Performance. Pharm. Res. 2020, 37, 101. [Google Scholar] [CrossRef]
  207. Nualnuk, P.; Sukpancharoen, S.; Srinophakun, T.R. Coupled Eulerian-Lagrangian Simulation of Spray-Drying Skimmed Milk Powder. Agric. Nat. Resour. 2024, 58, 33–42. [Google Scholar] [CrossRef]
  208. Kajishima, T.; Taira, K. Computational Fluid Dynamics; Springer International Publishing: Cham, Switzerland, 2017; ISBN 978-3-319-45302-6. [Google Scholar]
  209. Jaskulski, M.; Wawrzyniak, P.; Zbiciński, I. CFD Simulations of Droplet and Particle Agglomeration in an Industrial Counter-Current Spray Dryer. Adv. Powder Technol. 2018, 29, 1724–1733. [Google Scholar] [CrossRef]
  210. Anandharamakrishnan, C. Computational Fluid Dynamics Applications in Food Processing. In Computational Fluid Dynamics Applications in Food Processing; Springer New York: New York, NY, 2013; pp. 1–9. ISBN 978-1-4614-7990-1. [Google Scholar]
  211. Ducept, F.; Sionneau, M.; Vasseur, J. Superheated Steam Dryer: Simulations and Experiments on Product Drying. Chem. Eng. J. 2002, 86, 75–83. [Google Scholar] [CrossRef]
  212. Buchholz, M.; Haus, J.; Pietsch-Braune, S.; Kleine Jäger, F.; Heinrich, S. CFD-Aided Population Balance Modeling of a Spray Drying Process. Adv. Powder Technol. 2022, 33, 103636. [Google Scholar] [CrossRef]
  213. Mezhericher, M.; Levy, A.; Borde, I. Modelling of Particle Breakage during Drying. Chem. Eng. Process. Process Intensif. 2008, 47, 1404–1411. [Google Scholar] [CrossRef]
  214. Singh, M.; Ranade, V.; Shardt, O.; Matsoukas, T. Challenges and Opportunities Concerning Numerical Solutions for Population Balances: A Critical Review. J. Phys. A Math. Theor. 2022, 55, 383002. [Google Scholar] [CrossRef]
  215. Tang, Y.; Liu, C.; Zhang, X. Single Image Super-Resolution Using Wasserstein Generative Adversarial Network with Gradient Penalty. Pattern Recognit. Lett. 2022, 163, 32–39. [Google Scholar] [CrossRef]
  216. Bello, E.T.; Nasir, A.; Abdulkareem, A.S.; Muhammadu, M.M.; Musa, N.A.; Tijani, J.O.; Bada, S.O. Nano-Coolants for Thermal Enhancement in Heat Exchangers: A Review of Prospects, Challenges and Applications. Next Nanotechnol. 2025, 8, 100214. [Google Scholar] [CrossRef]
  217. Shrivastava, V.; Singh, P.; Thakur, V.K.; Abdelkader, T.K.; Yadav, A.S. Advancements in Computational Tools for Indirect Solar Drying Systems: A Comprehensive Review. Sustain. Energy Technol. Assess. 2025, 82, 104544. [Google Scholar] [CrossRef]
  218. Daskiran, C.; Riglin, J.; Schleicher, W.; Oztekin, A. Transient Analysis of Micro-Hydrokinetic Turbines for River Applications. Ocean Eng. 2017, 129, 291–300. [Google Scholar] [CrossRef]
  219. Lysenko, D.A. Free Stream Turbulence Intensity Effects on the Flow over a Circular Cylinder at Re = 3900: Bifurcation, Attractors and Lyapunov Metric. Ocean Eng. 2023, 287, 115787. [Google Scholar] [CrossRef]
  220. Kuang, L.; Katsuchi, H.; Zhou, D.; Chen, Y.; Han, Z.; Zhang, K.; Wang, J.; Bao, Y.; Cao, Y.; Liu, Y. Strategy for Mitigating Wake Interference Between Offshore Vertical-Axis Wind Turbines: Evaluation of Vertically Staggered Arrangement. Appl. Energy 2023, 351, 121850. [Google Scholar] [CrossRef]
  221. Román-Roldán, N.I.; López-Ortiz, A.; Ituna-Yudonago, J.F.; Nair, P.K.; Rodríguez-Ramírez, J.; Sandoval-Torres, S.; Martynenko, A. A Current Review: Engineering Design of Greenhouse Solar Dryers Exploring Novel Approaches. Sustain. Energy Technol. Assess. 2025, 73, 104137. [Google Scholar] [CrossRef]
  222. Hong, F.; Tian, H.; Yuan, X.; Liu, S.; Peng, Q.; Shi, Y.; Jin, L.; Ye, L.; Jia, J.; Ying, D.; et al. CFD-Assisted Modeling of the Hydrodynamic Cavitation Reactors for Wastewater Treatment—A Review. J. Environ. Manag. 2022, 321, 115982. [Google Scholar] [CrossRef] [PubMed]
  223. Jaskulski, M.; Wawrzyniak, P. Simulation and Design of Spray Driers. In Spray Drying for the Food Industry: Unit Operations and Processing Equipment in the Food Industry; Woodhead Publishing: Cambridge, UK, 2024; pp. 29–56. ISBN 9780128197998. [Google Scholar]
  224. Barbosa, D.V.E.; Santos, A.L.G.; dos Santos, E.D.; Souza, J.A. Overtopping Device Numerical Study: Openfoam Solution Verification and Evaluation of Curved Ramps Performances. Int. J. Heat Mass Transf. 2019, 131, 411–423. [Google Scholar] [CrossRef]
  225. López, A.; Nicholls, W.; Stickland, M.T.; Dempster, W.M. CFD Study of Jet Impingement Test Erosion Using Ansys Fluent® and OpenFOAM®. Comput. Phys. Commun. 2015, 197, 88–95. [Google Scholar] [CrossRef]
  226. Windt, C.; Davidson, J.; Ringwood, J. V High-Fidelity Numerical Modelling of Ocean Wave Energy Systems: A Review of Computational Fluid Dynamics-Based Numerical Wave Tanks. Renew. Sustain. Energy Rev. 2018, 93, 610–630. [Google Scholar] [CrossRef]
  227. Xuanyou, L. An Investigation of Error Sources in Computational Fluid Dynamics Modelling of a Co-Current Spray Dryer. Chin. J. Chem. Eng. 2004, 12, 756–761. [Google Scholar]
  228. Jubaer, H.; Afshar, S.; Xiao, J.; Chen, X.D.; Selomulya, C.; Woo, M.W. On the Importance of Droplet Shrinkage in CFD-Modeling of Spray Drying. Dry. Technol. 2018, 36, 1785–1801. [Google Scholar] [CrossRef]
  229. Mujumdar, A.S.; Huang, L.-X.; Chen, X.D. An Overview of the Recent Advances in Spray-Drying. Dairy Sci. Technol. 2010, 90, 211–224. [Google Scholar] [CrossRef]
  230. Hussain, F.; Jaskulski, M.; Piatkowski, M.; Tsotsas, E. CFD Simulation of Agglomeration and Coalescence in Spray Dryer. Chem. Eng. Sci. 2022, 247, 117064. [Google Scholar] [CrossRef]
  231. Jubaer, H. Computational Fluid Dynamics Modelling of Agglomeration during Spray Drying. Ph.D. Thesis, Monash University, Clayton, VIC, Australia, 2020. [Google Scholar]
  232. Ali, M.; Mahmud, T.; Heggs, P.; Ghadiri, M.; Bayly, A.; Crosby, M.; Ahmadian, H.; Martindejuan, L.; Alam, Z. Residence Time Distribution of Glass Ballotini in Isothermal Swirling Flows in a Counter-Current Spray Drying Tower. Powder Technol. 2017, 305, 809–815. [Google Scholar] [CrossRef]
  233. Sefidan, A.M.; Sellier, M.; Hewett, J.N.; Abdollahi, A.; Willmott, G.R.; Becker, S.M. Numerical Model to Study the Statistics of Whole Milk Spray Drying. Powder Technol. 2022, 411, 117923. [Google Scholar] [CrossRef]
  234. Breinlinger, T.; Hashibon, A.; Kraft, T. Simulation of the Influence of Surface Tension on Granule Morphology during Spray Drying Using a Simple Capillary Force Model. Powder Technol. 2015, 283, 1–8. [Google Scholar] [CrossRef]
  235. Kloss, C.; Goniva, C.; Hager, A.; Amberger, S.; Pirker, S. Models, Algorithms and Validation for Opensource DEM and CFD-DEM. Prog. Comput. Fluid Dyn. 2012, 12, 140–152. [Google Scholar] [CrossRef]
  236. Buchholz, M.; Weis, D.; Togni, R.; Goniva, C.; Heinrich, S. Describing the Drying and Solidification Behavior of Single Suspension Droplets Using a Novel Unresolved CFD-DEM Simulation Approach. Processes 2024, 12, 480. [Google Scholar] [CrossRef]
  237. Mezhericher, M.; Levy, A.; Borde, I. Multi-Scale Multiphase Modeling of Transport Phenomena in Spray-Drying Processes. Dry. Technol. 2015, 33, 2–23. [Google Scholar] [CrossRef]
  238. Xiao, J.; Chen, L.; Wu, W.D.; Chen, X.D. Multiscale Modeling for Nanoscale Surface Composition of Spray-Dried Powders: The Effect of Initial Droplet Size. Dry. Technol. 2016, 34, 1063–1072. [Google Scholar] [CrossRef]
  239. Sper de Almeida, E. Reducing Time Delays in Computing Numerical Weather Models at Regional and Local Levels: A Grid-Based Approach. Int. J. Grid Comput. Appl. 2012, 3, 1. [Google Scholar] [CrossRef]
  240. Rodrigues, A.E.; Paluch, A.S.; Andrés Gutiérrez Suárez, J.; Humberto, C.; Urueña, G.; Mejía, A.G. Adaptive Mesh Refinement Strategies for Cost-Effective Eddy-Resolving Transient Simulations of Spray Dryers. ChemEngineering 2023, 7, 100. [Google Scholar] [CrossRef]
  241. Anusha, P.; Kumar, S.S.; Pathipati, C. A Review of Machine Learning Strategies for Enhancing Efficiency and Innovation in Real-World Engineering Applications. Int. J. Sci. Res. Eng. Manag. 2024, 8, 1–5. [Google Scholar] [CrossRef]
  242. Jhaveri, R.H.; Revathi, A.; Ramana, K.; Raut, R.; Dhanaraj, R.K. A Review on Machine Learning Strategies for Real-World Engineering Applications. Mob. Inf. Syst. 2022, 2022, 1833507. [Google Scholar] [CrossRef]
  243. Thai, H.-T. Machine Learning for Structural Engineering: A State-of-the-Art Review. Structures 2022, 38, 448–491. [Google Scholar] [CrossRef]
  244. Zhong, S.; Zhang, K.; Bagheri, M.; Burken, J.G.; Gu, A.; Li, B.; Ma, X.; Marrone, B.L.; Ren, Z.J.; Schrier, J.; et al. Machine Learning: New Ideas and Tools in Environmental Science and Engineering. Environ. Sci. Technol. 2021, 55, 12741–12754. [Google Scholar] [CrossRef] [PubMed]
  245. Çınar, Z.M.; Abdussalam Nuhu, A.; Zeeshan, Q.; Korhan, O.; Asmael, M.; Safaei, B. Machine Learning in Predictive Maintenance towards Sustainable Smart Manufacturing in Industry 4.0. Sustainability 2020, 12, 8211. [Google Scholar] [CrossRef]
  246. Brunton, S.L.; Noack, B.R.; Koumoutsakos, P. Machine Learning for Fluid Mechanics. Annu. Rev. Fluid Mech. 2020, 52, 477–508. [Google Scholar] [CrossRef]
  247. Frank, M.; Drikakis, D.; Charissis, V. Machine-Learning Methods for Computational Science and Engineering. Computation 2020, 8, 15. [Google Scholar] [CrossRef]
  248. Lo, Y.; Rensi, S.E.; Torng, W.; Altman, R.B. Machine Learning in Chemoinformatics and Drug Discovery. Drug Discov. Today 2018, 23, 1538–1546. [Google Scholar] [CrossRef]
  249. Alam, G.; Ihsanullah, I.; Naushad, M.; Sillanpää, M. Applications of Artificial Intelligence in Water Treatment for Optimization and Automation of Adsorption Processes: Recent Advances and Prospects. Chem. Eng. J. 2022, 427, 130011. [Google Scholar] [CrossRef]
  250. Woldaregay, A.Z.; Årsand, E.; Walderhaug, S.; Albers, D.; Mamykina, L.; Botsis, T.; Hartvigsen, G. Data-Driven Modeling and Prediction of Blood Glucose Dynamics: Machine Learning Applications in Type 1 Diabetes. Artif. Intell. Med. 2019, 98, 109–134. [Google Scholar] [CrossRef]
  251. Yang, H.S.; Rhoads, D.D.; Sepulveda, J.; Zang, C.; Chadburn, A.; Wang, F. Building the Model: Challenges and Considerations of Developing and Implementing Machine Learning Tools for Clinical Laboratory Medicine Practice. Arch. Pathol. Lab. Med. 2023, 147, 826–836. [Google Scholar] [CrossRef]
  252. Wang, F.; Casalino, L.P.; Khullar, D. Deep Learning in Medicine—Promise, Progress, and Challenges. JAMA Intern. Med. 2019, 179, 293–294. [Google Scholar] [CrossRef] [PubMed]
  253. Thornton, C.; Hutter, F.; Hoos, H.H.; Leyton-Brown, K. Auto-WEKA: Combined Selection and Hyperparameter Optimization of Classification Algorithms. In Proceedings of the 19th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, Chicago, OH, USA, 11–14 August 2013; Association for Computing Machinery: New York, NY, USA, 2013; Volume Part F128815, pp. 847–855. [Google Scholar]
  254. Zheng, J.; Ye, L.; Ge, Z. Laplacian Regularization of Linear Regression Model for Semi-Supervised Industrial Soft Sensor Development. Expert. Syst. Appl. 2024, 254, 124459. [Google Scholar] [CrossRef]
  255. Bystritskaya, E.V.; Pomerantsev, A.L.; Rodionova, O.Y. Non-Linear Regression Analysis: New Approach to Traditional Implementations. J. Chemom. 2000, 14, 667–692. [Google Scholar] [CrossRef]
  256. Bharadwaj; Prakash, K.B.; Kanagachidambaresan, G.R. Pattern Recognition and Machine Learning BT. In Programming with TensorFlow: Solution for Edge Computing Applications; Prakash, K.B., Kanagachidambaresan, G.R., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 105–144. ISBN 978-3-030-57077-4. [Google Scholar]
  257. Zhang, X. Support Vector Machines BT. In Encyclopedia of Machine Learning and Data Mining; Sammut, C., Webb, G.I., Eds.; Springer US: Boston, MA, USA, 2017; pp. 1214–1220. ISBN 978-1-4899-7687-1. [Google Scholar]
  258. Rodríguez-Pérez, R.; Bajorath, J. Evolution of Support Vector Machine and Regression Modeling in Chemoinformatics and Drug Discovery. J. Comput. Aided Mol. Des. 2022, 36, 355–362. [Google Scholar] [CrossRef]
  259. Saravanan, K.; Prakash, R.B.; Balakrishnan, C.; Kumar, G.V.P.; Siva Subramanian, R.; Anita, M. Support Vector Machines: Unveiling the Power and Versatility of SVMs in Modern Machine Learning. In Proceedings of the 3rd International Conference on Innovative Mechanisms for Industry Applications, ICIMIA 2023, Bengaluru, India, 21–23 December 2023; Institute of Electrical and Electronics Engineers Inc.: New York, NY, USA, 2023; pp. 680–687. [Google Scholar]
  260. Fradi, A.; Tran, T.T.; Samir, C. Decomposed Gaussian Processes for Efficient Regression Models with Low Complexity. Entropy 2025, 27, 393. [Google Scholar] [CrossRef]
  261. Zimmermann, A. Ensemble-Trees: Leveraging Ensemble Power Inside Decision Trees. In Proceedings of the Discovery Science, Budapest, Hungary, 13–16 October 2008; Jean-Fran, J.-F., Berthold, M.R., Horváth, T., Eds.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 76–87. [Google Scholar]
  262. Cheng, W.X.; Katuwal, R.; Suganthan, P.N.; Qiu, X. A Heterogeneous Ensemble of Trees. In Proceedings of the IEEE Symposium Series on Computational Intelligence (SSCI), Honolulu, HI, USA, 27 November–1 December 2017; pp. 1–6. [Google Scholar]
  263. Singh, U.; Rizwan, M.; Alaraj, M.; Alsaidan, I. A Machine Learning-Based Gradient Boosting Regression Approach for Wind Power Production Forecasting: A Step towards Smart Grid Environments. Energies 2021, 14, 5196. [Google Scholar] [CrossRef]
  264. Taylan, O.A. Artificial Neural Network Models in Prediction of the Moisture Content of a Spray Drying Process. J. Korean Ceram. Soc. 2004, 41, 353–358. [Google Scholar] [CrossRef]
  265. Abdolrasol, M.G.M.; Suhail Hussain, S.M.; Ustun, T.S.; Sarker, M.R.; Hannan, M.A.; Mohamed, R.; Ali, J.A.; Mekhilef, S.; Milad, A. Artificial Neural Networks Based Optimization Techniques: A Review. Electronics 2021, 10, 2689. [Google Scholar] [CrossRef]
  266. Achanta, A.S.; Kowalski, J.; Rhodes, C.T. Artificial Neural Networks: Implications for Pharmaceutical Sciences. Drug Dev. Ind. Pharm. 1995, 21, 119–155. [Google Scholar] [CrossRef]
  267. Beresford, R. Basic Concepts of Artificial Neural Network (ANN) Modeling and Its Application in Pharmaceutical Research. J. Pharm. Biomed. Anal. 2000, 22, 717–727. [Google Scholar] [CrossRef]
  268. Rumelhart, D.E.; Widrow, B.; Lehr, M.A. The Basic Ideas in Neural Networks. Commun. ACM 1994, 37, 87–92. [Google Scholar] [CrossRef]
  269. Jiang, J.; Peng, H.H.; Yang, Z.; Ma, X.; Sahakijpijarn, S.; Moon, C.; Ouyang, D.; Williams, R.O. The Applications of Machine Learning (ML) in Designing Dry Powder for Inhalation by Using Thin-Film-Freezing Technology. Int. J. Pharm. 2022, 626, 122179. [Google Scholar] [CrossRef] [PubMed]
  270. Lisboa, H.M.; Lúcio, A.; Andrade, R.; Sarinho, A.M.; Lima, J.; Batista, L.; Costa, M.E.; Nascimento, A.; Pasquali, M.B. Leveraging Quantitative Structure-Activity Relationships to Design and Optimize Wall Material Formulations for Antioxidant Encapsulation via Spray Drying. Food Chem. Adv. 2025, 7, 100948. [Google Scholar] [CrossRef]
  271. Mallamma, T.; Raghavendra Naveen, N.; Goudanavar, P. Spray Dried Linezolid-Loaded Inhalable Chitosan Microparticles for Tuberculosis: A QbD-Driven Approach to Enhance Pulmonary Delivery. Biomed. Mater. Devices 2025, 3, 1351–1363. [Google Scholar] [CrossRef]
  272. Bao, Z. Towards Data-Driven Development of Advanced Drug Formulations Leveraging Machine Learning and Experimental Automation. Ph.D. Thesis, University of Toronto, Toronto, ON, Canada, 2024. [Google Scholar]
  273. Li, Y.; Dong, X.; Cao, G.; Guo, Y.; Wang, Z.; Yang, X.; Han, D.; Meng, Z. Identifying Baicalin Concentration in Scutellaria Spray Drying Powder With Disturbed Terahertz Spectra Based on Gaussian Mixture Model. J. Anal. Methods Chem. 2024, 2024, 3858763. [Google Scholar] [CrossRef]
  274. Moore, M.D.; Wildfong, P.L.D. Informatics Calibration of a Molecular Descriptors Database to Predict Solid Dispersion Potential of Small Molecule Organic Solids. Int. J. Pharm. 2011, 418, 217–226. [Google Scholar] [CrossRef]
  275. Nurzyńska, K.; Booth, J.; Roberts, C.J.; McCabe, J.; Dryden, I.; Fischer, P.M. Long-Term Amorphous Drug Stability Predictions Using Easily Calculated, Predicted, and Measured Parameters. Mol. Pharm. 2015, 12, 3389–3398. [Google Scholar] [CrossRef]
  276. Kapourani, A.; Eleftheriadou, K.; Kontogiannopoulos, K.N.; Barmpalexis, P. Evaluation of Rivaroxaban Amorphous Solid Dispersions Physical Stability via Molecular Mobility Studies and Molecular Simulations. Eur. J. Pharm. Sci. 2021, 157, 105642. [Google Scholar] [CrossRef]
  277. Medarević, D.P.; Kachrimanis, K.; Mitrić, M.; Djuriš, J.; Djurić, Z.; Ibrić, S. Dissolution Rate Enhancement and Physicochemical Characterization of Carbamazepine-Poloxamer Solid Dispersions. Pharm. Dev. Technol. 2016, 21, 268–276. [Google Scholar] [CrossRef]
  278. Patel, A.D.; Agrawal, A.; Dave, R.H. European Journal of Pharmaceutics and Biopharmaceutics Investigation of the Effects of Process Variables on Derived Properties of Spray Dried Solid-Dispersions Using Polymer Based Response Surface Model and Ensemble Artificial Neural Network Models. Eur. J. Pharm. Biopharm. 2014, 86, 404–417. [Google Scholar] [CrossRef]
  279. Muñiz Castro, B.; Elbadawi, M.; Ong, J.J.; Pollard, T.; Song, Z.; Gaisford, S.; Pérez, G.; Basit, A.W.; Cabalar, P.; Goyanes, A. Machine Learning Predicts 3D Printing Performance of over 900 Drug Delivery Systems. J. Control. Release 2021, 337, 530–545. [Google Scholar] [CrossRef] [PubMed]
  280. Abbas, A.; Farizhandi, K.; Alishiri, M.; Lau, R. Machine Learning Approach for Carrier Surface Design in Carrier-Based Dry Powder Inhalation. Comput. Chem. Eng. 2021, 151, 107367. [Google Scholar] [CrossRef]
  281. Ma, X.; Kittikunakorn, N.; Sorman, B.; Xi, H.; Chen, A.; Marsh, M.; Mongeau, A.; Piché, N.; Williams, R.O.; Skomski, D. Application of Deep Learning Convolutional Neural Networks for Internal Tablet Defect Detection: High Accuracy, Throughput, and Adaptability. J. Pharm. Sci. 2020, 109, 1547–1557. [Google Scholar] [CrossRef] [PubMed]
  282. Laitinen, R.; Löbmann, K.; Strachan, C.J.; Grohganz, H.; Rades, T. Emerging Trends in the Stabilization of Amorphous Drugs. Int. J. Pharm. 2013, 453, 65–79. [Google Scholar] [CrossRef]
  283. Bock, S.; Rades, T.; Rantanen, J.; Scherließ, R.; Cui, P.; Yu, Y.; Zhao, J.; Miao, P.; Xue, Q.; Liu, C.; et al. Combining Machine Learning and Molecular Simulations to Predict the Stability of Amorphous Drugs. J. Chem. Phys. 2023, 159, 113246. [Google Scholar] [CrossRef]
  284. Hoseini, B.; Jaafari, M.R.; Golabpour, A.; Momtazi-Borojeni, A.A.; Karimi, M.; Eslami, S. Application of Ensemble Machine Learning Approach to Assess the Factors Affecting Size and Polydispersity Index of Liposomal Nanoparticles. Sci. Rep. 2023, 13, 18012. [Google Scholar] [CrossRef]
  285. Galata, D.L.; Könyves, Z.; Nagy, B.; Novák, M.; Mészáros, L.A.; Szabó, E.; Farkas, A.; Marosi, G.; Nagy, Z.K. Real-Time Release Testing of Dissolution Based on Surrogate Models Developed by Machine Learning Algorithms Using NIR Spectra, Compression Force and Particle Size Distribution as Input Data. Int. J. Pharm. 2021, 597, 120338. [Google Scholar] [CrossRef]
  286. Nespi, M.; Kuhn, R.; Yen, C.W.; Leung, D. Optimization of Spray-Drying Parameters for Formulation Development at Preclinical Scale. AAPS PharmSciTech 2022, 23, 28. [Google Scholar] [CrossRef]
  287. Patel, S.; Patel, M.; Kulkarni, M.; Patel, M.S. DE-INTERACT: A Machine-Learning-Based Predictive Tool for the Drug-Excipient Interaction Study during Product Development—Validation through Paracetamol and Vanillin as a Case Study. Int. J. Pharm. 2023, 637, 122839. [Google Scholar] [CrossRef]
  288. Han, R.; Xiong, H.; Ye, Z.; Yang, Y.; Huang, T.; Jing, Q.; Lu, J.; Pan, H.; Ren, F.; Ouyang, D. Predicting Physical Stability of Solid Dispersions by Machine Learning Techniques. J. Control. Release 2019, 311–312, 16–25. [Google Scholar] [CrossRef]
  289. Dong, J.; Gao, H.; Ouyang, D. PharmSD: A Novel AI-Based Computational Platform for Solid Dispersion Formulation Design. Int. J. Pharm. 2021, 604, 120705. [Google Scholar] [CrossRef] [PubMed]
  290. Wagh, N.; Agashe, S.D. Monitoring Exhaust Air Temperature and Detecting Faults during Milk Spray Drying Using Data-Driven Framework. Dry. Technol. 2023, 41, 1991–2011. [Google Scholar] [CrossRef]
  291. Simões, M.F.; Silva, G.; Pinto, A.C.; Fonseca, M.; Silva, N.E.; A, B.I.F.S. European Journal of Pharmaceutics and Biopharmaceutics Artificial Neural Networks Applied to Quality-by-Design: From Formulation Development to Clinical Outcome. Eur. J. Pharm. Biopharm. 2020, 152, 282–295. [Google Scholar] [CrossRef] [PubMed]
  292. Lou, H.; Chung, J.I.; Kiang, Y.H.; Xiao, L.Y.; Hageman, M.J. The Application of Machine Learning Algorithms in Understanding the Effect of Core/Shell Technique on Improving Powder Compactability. Int. J. Pharm. 2019, 555, 368–379. [Google Scholar] [CrossRef]
  293. Miletić, T.; Ibrić, S.; Đurić, Z. Combined Application of Experimental Design and Artificial Neural Networks in Modeling and Characterization of Spray Drying Drug: Cyclodextrin Complexes. Dry. Technol. 2014, 32, 167–179. [Google Scholar] [CrossRef]
  294. Fiedler, D.; Fink, E.; Aigner, I.; Leitinger, G.; Keller, W.; Roblegg, E.; Khinast, J.G. A Multi-Step Machine Learning Approach for Accelerating QbD-Based Process Development of Protein Spray Drying. Int. J. Pharm. 2023, 642, 123133. [Google Scholar] [CrossRef]
  295. Milanesi, A.; Rizzuto, F.; Rinaldi, M.; Foglio Bonda, A.; Segale, L.; Giovannelli, L. Thermodynamic Balance vs. Computational Fluid Dynamics Approach for the Outlet Temperature Estimation of a Benchtop Spray Dryer. Pharmaceutics 2022, 14, 296. [Google Scholar] [CrossRef]
  296. Tathawadekar, N.; Doan, N.A.K.; Silva, C.F.; Thuerey, N. Modeling of the Nonlinear Flame Response of a Bunsen-Type Flame via Multi-Layer Perceptron. Proc. Combust. Inst. 2021, 38, 6261–6269. [Google Scholar] [CrossRef]
  297. Li, F.; Ma, G.; Ju, C.; Chen, S.; Huang, W. Data-Driven Forecasting Framework for Daily Reservoir Inflow Time Series Considering the Flood Peaks Based on Multi-Head Attention Mechanism. J. Hydrol. 2024, 645, 132197. [Google Scholar] [CrossRef]
  298. Xiao, J.; Li, Y.; George, O.A.; Li, Z.; Yang, S.; Woo, M.W.; Wu, W.D.; Chen, X.D. Numerical Investigation of Droplet Pre-Dispersion in a Monodisperse Droplet Spray Dryer. Particuology 2018, 38, 44–60. [Google Scholar] [CrossRef]
  299. Duarte, V.; Duarte, D.; Fonseca, J.; Montecinos, A. Benchmarking Machine-Learning Software and Hardware for Quantitative Economics. J. Econ. Dyn. Control 2020, 111, 103796. [Google Scholar] [CrossRef]
  300. Lin, J.-K.; Nurazaq, W.A.; Wang, W.-C.; Lee, C.-Y.; Poon, H.M.; Gan, S.; Duong, V.D.; Prapainainar, P. Unsteady Spray Dynamics of Hydro-Processed Renewable Diesel: Influence of Thermo-Physical Properties under High-Pressure Injection. J. Energy Inst. 2025, 122, 102222. [Google Scholar] [CrossRef]
  301. Martynenko, A.; Misra, N.N. Machine Learning in Drying. Dry. Technol. 2020, 38, 596–609. [Google Scholar] [CrossRef]
  302. Ajmal, C.S.; Yerram, S.; Abishek, V.; Nizam, V.P.M.; Aglave, G.; Patnam, J.D.; Raghuvanshi, R.S.; Srivastava, S. Innovative Approaches in Regulatory Affairs: Leveraging Artificial Intelligence and Machine Learning for Efficient Compliance and Decision-Making. AAPS J. 2025, 27, 22. [Google Scholar] [CrossRef] [PubMed]
  303. Mirakhori, F.; Niazi, S.K. Harnessing the AI/ML in Drug and Biological Products Discovery and Development: The Regulatory Perspective. Pharmaceuticals 2025, 18, 47. [Google Scholar] [CrossRef] [PubMed]
  304. Niazi, S.K. Regulatory Perspectives for AI/ML Implementation in Pharmaceutical GMP Environments. Pharmaceuticals 2025, 18, 901. [Google Scholar] [CrossRef] [PubMed]
  305. Pantanowitz, L.; Hanna, M.; Pantanowitz, J.; Lennerz, J.; Henricks, W.H.; Shen, P.; Quinn, B.; Bennet, S.; Rashidi, H.H. Regulatory Aspects of Artificial Intelligence and Machine Learning. Mod. Pathol. 2024, 37, 100609. [Google Scholar] [CrossRef] [PubMed]
  306. Kavasidis, I.; Lallas, E.; Leligkou, H.C.; Oikonomidis, G.; Karydas, D.; Gerogiannis, V.C.; Karageorgos, A. Deep Transformers for Computing and Predicting ALCOA+Data Integrity Compliance in the Pharmaceutical Industry. Appl. Sci. 2023, 13, 7616. [Google Scholar] [CrossRef]
  307. Salih, A.M.; Raisi-Estabragh, Z.; Galazzo, I.B.; Radeva, P.; Petersen, S.E.; Lekadir, K.; Menegaz, G. A Perspective on Explainable Artificial Intelligence Methods: SHAP and LIME. Adv. Intell. Syst. 2025, 7, 2400304. [Google Scholar] [CrossRef]
  308. Mirbabaie, M.; Stieglitz, S.; Frick, N.R.J. Artificial Intelligence in Disease Diagnostics: A Critical Review and Classification on the Current State of Research Guiding Future Direction. Health Technol. 2021, 11, 693–731. [Google Scholar] [CrossRef]
  309. Manzano, T.; Whitford, W. AI Applications for Multivariate Control in Drug Manufacturing. In A Handbook of Artificial Intelligence in Drug Delivery; Philip, A., Shahiwala, A., Rashid, M., Faiyazuddin, M., Eds.; Academic Press: London, UK, 2023; pp. 55–82. ISBN 9780323899253. [Google Scholar]
  310. Parikh, R.B.; Helmchen, L.A. Paying for Artificial Intelligence in Medicine. NPJ Digit. Med. 2022, 5, 63. [Google Scholar] [CrossRef]
  311. Wei, W.; Song, C.; Qi, C.; Zhang, X.; Zhang, X.; Pu, R.; Ao, Y. Visual Analysis of Drug Research and Development Based on Artificial Intelligence. J. Holist. Integr. Pharm. 2024, 5, 323–332. [Google Scholar] [CrossRef]
  312. Liang, N.L.; Guffey, M.; Jeon, B.; Vorp, D.A.; Chung, T.K. The Evolving Regulatory Landscape for Artificial Intelligence-Enabled Devices in the United States. JVS-Vasc. Insights 2025, 3, 100290. [Google Scholar] [CrossRef]
  313. Duong-Trung, N.; Born, S.; Kim, J.W.; Schermeyer, M.-T.; Paulick, K.; Borisyak, M.; Cruz-Bournazou, M.N.; Werner, T.; Scholz, R.; Schmidt-Thieme, L.; et al. When Bioprocess Engineering Meets Machine Learning: A Survey from the Perspective of Automated Bioprocess Development. Biochem. Eng. J. 2023, 190, 108764. [Google Scholar] [CrossRef]
  314. Guerra-Manzanares, A.; Caprolu, M.; Di Pietro, R. A Comprehensive Review on Machine Learning-Based VPN Detection: Scenarios, Methods, and Open Challenges. Comput. Sci. Rev. 2025, 58, 100781. [Google Scholar] [CrossRef]
  315. James, A.; Vajrashree, P.S.; Meghana, G.S.; Sastri, K.T.; Akhila, A.R.; Gowda, C.U.; Balamuralidhara , V. Beyond the Blood-Brain Barrier: Intranasal Vaccines in Cancer Immunotherapy. Crit. Rev. Oncol. Hematol. 2025, 216, 104938. [Google Scholar] [CrossRef] [PubMed]
  316. Suffian, M.; Kuhl, U.; Bogliolo, A.; Alonso-Moral, J.M. The Role of User Feedback in Enhancing Understanding and Trust in Counterfactual Explanations for Explainable AI. Int. J. Hum. Comput. Stud. 2025, 199, 103484. [Google Scholar] [CrossRef]
  317. Benito, D.; Quintero, C.; Aguilar, J.; Ramírez, J.M.; Fernández-Anta, A. Explainability Analysis: An in-Depth Comparison between Fuzzy Cognitive Maps and LAMDA. Appl. Soft Comput. 2024, 164, 111940. [Google Scholar] [CrossRef]
  318. Martinis, M.C.; Zucco, C. Artificial Intelligence and Explainability. In Encyclopedia of Bioinformatics and Computational Biology; Ranganathan, S., Cannataro, M., Khan, A., Eds.; Elsevier: Oxford, UK, 2025; pp. 393–398. ISBN 978-0-323-95503-4. [Google Scholar]
  319. Fujiwara, K. Knowledge Distillation with Resampling for Imbalanced Data Classification: Enhancing Predictive Performance and Explainability Stability. Results Eng. 2024, 24, 103406. [Google Scholar] [CrossRef]
  320. He, H.-H.; Li, W.; Zhu, Y.; Tao, Z.; Zhao, J. Machine Learning Analysis for Condensation Flow Heat Transfer in Mini/Micro-Channels. Int. J. Heat Mass Transf. 2026, 255, 127775. [Google Scholar] [CrossRef]
  321. Procopio, A.; Cesarelli, G.; Donisi, L.; Merola, A.; Amato, F.; Cosentino, C. Combined Mechanistic Modeling and Machine-Learning Approaches in Systems Biology—A Systematic Literature Review. Comput. Methods Programs Biomed. 2023, 240, 107681. [Google Scholar] [CrossRef] [PubMed]
  322. McBride, K.; Sundmacher, K. Overview of Surrogate Modeling in Chemical Process Engineering. Chem. Ingen. Tech. 2019, 91, 228–239. [Google Scholar] [CrossRef]
  323. Mihajlovic, T.; Svetlana, I.; Mladenovic, A. Application of Design of Experiments and Multilayer Perceptron Neural Network in Optimization of the Spray-Drying Process. Dry. Technol. 2011, 29, 1638–1647. [Google Scholar] [CrossRef]
  324. Merkelbach, K.; Schweidtmann, A.M.; Müller, Y.; Schwoebel, P.; Mhamdi, A.; Mitsos, A.; Schuppert, A.; Mrziglod, T.; Schneckener, S. HybridML: Open Source Platform for Hybrid Modeling. Comput. Chem. Eng. 2022, 160, 107736. [Google Scholar] [CrossRef]
  325. Sivasankaran, S.; Jonnalagadda, S. Levonorgestrel Loaded Biodegradable Microparticles for Injectable Contraception: Preparation, Characterization and Modelling of Drug Release. Int. J. Pharm. 2022, 624, 121994. [Google Scholar] [CrossRef]
  326. Zhou, T.; Song, Z.; Sundmacher, K. Big Data Creates New Opportunities for Materials Research: A Review on Methods and Applications of Machine Learning for Materials Design. Engineering 2019, 5, 1017–1026. [Google Scholar] [CrossRef]
  327. Zhang, C.; Gao, G.; Guo, Y.; Liu, Y.; Liu, Y.; Wu, G. Structural Optimization Model of Oil-Natural Air-Natural Transformer Radiator Based on Data-Model Hybrid-Driven. Appl. Therm. Eng. 2025, 260, 125016. [Google Scholar] [CrossRef]
  328. Hua, P.; Huang, H.; He, Y.; Zhang, J.; Ying, G. Hydrology-Informed Machine Learning for Enhancing Surface Runoff Elucidation. Sci. Total Environ. 2025, 992, 179936. [Google Scholar] [CrossRef]
  329. Cuesta, J.; Leturiondo, U.; Vidal, Y.; Pozo, F. A Review of Prognostics and Health Management Techniques in Wind Energy. Reliab. Eng. Syst. Saf. 2025, 260, 111004. [Google Scholar] [CrossRef]
  330. Zhu, S.; Gutterman, C.; Montiel, A.D.; Yu, J.; Ruffini, M.; Zussman, G.; Kilper, D. Hybrid Machine Learning EDFA Model. In Proceedings of the 2020 Optical Fiber Communication Conference (OFC), San Diego, CA, USA, 8–12 March 2020; Volume Part F174-OFC 2020, pp. 1–3. [Google Scholar]
  331. Kapil, A.; Wombwell, C.; Kegel, L.L.; Hamlin, M.D. Model-Aided Process Development for Scalable Spray Drying of Sticky Substances. Front. Chem. Eng. 2025, 7, 1565916. [Google Scholar] [CrossRef]
  332. Cui, P.; Yu, Y.; Xue, Q.; Wu, Z.; Miao, K.; Liu, C.; Zhao, L.; Li, Z. Numerical Simulation and Optimization of Lonicerae Japonicae Flos Extract Spray Drying Process Based on Temperature Field Verification and Deep Reinforcement Learning. J. Food Eng. 2023, 345, 111425. [Google Scholar] [CrossRef]
  333. Schweidtmann, A.M.; Esche, E.; Fischer, A.; Kloft, M.; Repke, J.U.; Sager, S.; Mitsos, A. Machine Learning in Chemical Engineering: A Perspective. Chem. Ing. Tech. 2021, 93, 2029–2039. [Google Scholar] [CrossRef]
  334. Panjapornpon, C.; Bardeeniz, S.; Hussain, M.A. Deep Learning Approach for Energy Efficiency Prediction with Signal Monitoring Reliability for a Vinyl Chloride Monomer Process. Reliab. Eng. Syst. Saf. 2023, 231, 109008. [Google Scholar] [CrossRef]
  335. Wang, K.-J.; Lee, Y.-H.; Angelica, S. Digital Twin Design for Real-Time Monitoring—A Case Study of Die Cutting Machine. Int. J. Prod. Res. 2021, 59, 6471–6485. [Google Scholar] [CrossRef]
  336. Purlis, E. Digital Twin Methodology in Food Processing: Basic Concepts and Applications. Curr. Nutr. Rep. 2024, 13, 914–920. [Google Scholar] [CrossRef]
  337. Bardeeniz, S.; Panjapornpon, C.; Fongsamut, C.; Ngaotrakanwiwat, P.; Azlan Hussain, M. Digital Twin-Aided Transfer Learning for Energy Efficiency Optimization of Thermal Spray Dryers: Leveraging Shared Drying Characteristics across Chemicals with Limited Data. Appl. Therm. Eng. 2024, 242, 122431. [Google Scholar] [CrossRef]
  338. Bolbotinović, Ž.; Milić, S.D.; Janda, Ž.; Vukmirović, D. Ai-Powered Digital Twin in the Industrial IoT. Int. J. Electr. Power Energy Syst. 2025, 167, 110656. [Google Scholar] [CrossRef]
  339. Kumar, R.; Agrawal, N. Shaping the Future of Industry: Understanding the Dynamics of Industrial Digital Twins. Comput. Ind. Eng. 2024, 191, 110172. [Google Scholar] [CrossRef]
  340. Vashishth, T.K.; Sharma, V.; Sharma, K.K.; Kumar, B.; Chaudhary, S.; Panwar, R. Digital Twins Solutions for Smart Logistics and Transportation. In Digital Twins for Smart Cities and Villages; Iyer, S., Nayyar, A., Paul, A., Naved, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2025; pp. 353–376. ISBN 978-0-443-28884-5. [Google Scholar]
  341. Ahmad, A.; Shah, A.; Alnumay, W.S.; Adnan, M.; Anwer, S.; Zaman, Q.U. A Comprehensive Survey on the Convergence of Blockchain, Digital Twins, and Metaverse: Shaping the Future of Cybersecurity Frameworks. Comput. Electr. Eng. 2025, 126, 110486. [Google Scholar] [CrossRef]
  342. Chaiburi, C.; Yingngam, B. Digital Twin Technology Fundamentals. In Digital Twins for Smart Cities and Villages; Iyer, S., Nayyar, A., Paul, A., Naved, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2025; pp. 1–35. ISBN 978-0-443-28884-5. [Google Scholar]
  343. Mezouar, H.; El Afia, A. A Systematic Literature Review of Machine Learning Applications in Software Engineering; Springer: Cham, Switzerland, 2021; ISBN 978-3-031-07968-9. [Google Scholar]
  344. Polyzotis, N.; Roy, S.; Whang, S.E.; Zinkevich, M.A. Data Lifecycle Challenges in Production Machine Learning. ACM Sigmod Rec. 2018, 47, 17–28. [Google Scholar] [CrossRef]
  345. Yu, Y. Data-Driven Innovation or Innovation-Driven Data Generation? In Proceedings of the 2023 IEEE International Conference on Engineering, Technology and Innovation (ICE/ITMC), Edinburgh, UK, 19–22 June 2023; pp. 1–8. [Google Scholar]
  346. Rajkumar, R.R.; Lee, I.; Sha, L.R.; Stankovic, J.A. Cyber-Physical Systems: The Next Computing Revolution. In Proceedings of the 47th Design Automation Conference, Anaheim, CA, USA, 13–18 June 2010; pp. 731–736. [Google Scholar]
  347. Barocas, S.; Selbst, A.D. Big Data’s Disparate Impact. Calif. Law Rev. 2016, 104, 671. [Google Scholar] [CrossRef]
  348. Buolamwini, J.; Gebru, T. Gender Shades: Intersectional Accuracy Disparities in Commercial Gender Classification. In Proceedings of the 1st Conference on Fairness, Accountability and Transparency, New York, NY, USA, 23–24 February 2018; pp. 1–10. [Google Scholar]
  349. Poozesh, S.; Lu, K.; Marsac, P.J. On the Particle Formation in Spray Drying Process for Bio-Pharmaceutical Applications: Interrogating a New Model via Computational Fluid Dynamics. Int. J. Heat Mass Transf. 2018, 122, 863–876. [Google Scholar] [CrossRef]
  350. Matsunami, K.; Ryckaert, A.; Verrecas, H.; da Silva, A.F.T.; Anderson, A.; Wikström, H.; Dumarey, M.; Mann, J.; De Beer, T.; Vanhoorne, V.; et al. Comparative Analysis of Tablet Dissolution Behavior: Batch vs. Continuous Direct Compression. Int. J. Pharm. 2025, 675, 125498. [Google Scholar] [CrossRef] [PubMed]
  351. Lisboa, H.M.; Duarte, M.E.; Cavalcanti-Mata, M.E. Modeling of Food Drying Processes in Industrial Spray Dryers. Food Bioprod. Process. 2018, 107, 49–60. [Google Scholar] [CrossRef]
  352. Baassiri, M.; Ranade, V.; Padrela, L. CFD Modelling and Simulations of Atomization-Based Processes for Production of Drug Particles: A Review. Int. J. Pharm. 2025, 670, 125204. [Google Scholar] [CrossRef]
  353. Zhang, H.; Huang, J.; Zhang, X.; Wong, C.-N. Autonomous Optimization of Process Parameters and In-Situ Anomaly Detection in Aerosol Jet Printing by an Integrated Machine Learning Approach. Addit. Manuf. 2024, 86, 104208. [Google Scholar] [CrossRef]
  354. Mnasri, Z.; D’Andrea, A. Automatic Inventory of Archaeological Artifacts Based on Object Detection and Classification Using Deep and Transfer Learning. Digit. Appl. Archaeol. Cult. Herit. 2025, 39, e00458. [Google Scholar] [CrossRef]
  355. Lamouadene, H.; EL Kassaoui, M.; El Yadari, M.; El Kenz, A.; Benyoussef, A.; El Moutaouakil, A.; Mounkachi, O. Detection of COVID-19, Lung Opacity, and Viral Pneumonia via X-Ray Using Machine Learning and Deep Learning. Comput. Biol. Med. 2025, 191, 110131. [Google Scholar] [CrossRef]
  356. Jaskulski, M.; Atuonwu, J.C.; Tran, T.T.H.; Stapley, A.G.F.; Tsotsas, E. Predictive CFD Modeling of Whey Protein Denaturation in Skim Milk Spray Drying Powder Production. Adv. Powder Technol. 2017, 28, 3140–3147. [Google Scholar] [CrossRef]
Figure 1. Spray drying process flow diagram, reproduced from [51] with permission from Springer Nature, 2025.
Figure 1. Spray drying process flow diagram, reproduced from [51] with permission from Springer Nature, 2025.
Pharmaceutics 17 01605 g001
Figure 2. Evaporation and drying steps of a single droplet in a spray drying process: (A) Heat and mass transfer on the surface of the droplet; (B) Schematic representation of various drying stages [112].
Figure 2. Evaporation and drying steps of a single droplet in a spray drying process: (A) Heat and mass transfer on the surface of the droplet; (B) Schematic representation of various drying stages [112].
Pharmaceutics 17 01605 g002
Figure 3. Particle formation process: (A) Drying of droplets containing large and small solids volume fractions; (B) Influence of Peclet number on final particle morphology.
Figure 3. Particle formation process: (A) Drying of droplets containing large and small solids volume fractions; (B) Influence of Peclet number on final particle morphology.
Pharmaceutics 17 01605 g003
Figure 4. Effect of process parameters on moisture content of spray-dried powder: (A) Inlet temperature [144], (B) Feed flow rate [145].
Figure 4. Effect of process parameters on moisture content of spray-dried powder: (A) Inlet temperature [144], (B) Feed flow rate [145].
Pharmaceutics 17 01605 g004
Figure 5. SEM images of spray-dried powder at various outlet temperatures: (A) outlet drying temperature at 114 °C; (B) outlet drying temperature at 140 °C, reproduced from [150], with permission from Elsevier, 2025.
Figure 5. SEM images of spray-dried powder at various outlet temperatures: (A) outlet drying temperature at 114 °C; (B) outlet drying temperature at 140 °C, reproduced from [150], with permission from Elsevier, 2025.
Pharmaceutics 17 01605 g005
Figure 6. Shell formation predicted at solids volume fraction = 0.65, time = 60 s. by the model, reproduced from [179] with permission from Elsevier, 2025.
Figure 6. Shell formation predicted at solids volume fraction = 0.65, time = 60 s. by the model, reproduced from [179] with permission from Elsevier, 2025.
Pharmaceutics 17 01605 g006
Figure 8. Computational fluid dynamics (CFD) simulation workflow: (A) Stages involved in CFD Analysis; (B) Parameters obtained in the spray drying process using CFD simulation, adapted from [198], with permission from Elsevier, 2025.
Figure 8. Computational fluid dynamics (CFD) simulation workflow: (A) Stages involved in CFD Analysis; (B) Parameters obtained in the spray drying process using CFD simulation, adapted from [198], with permission from Elsevier, 2025.
Pharmaceutics 17 01605 g008
Figure 10. Machine learning workflow: (A) Steps involved in developing AI/ML-based models; (B) Different levels of complexity of ML models, ranging from simple linear regression models to more complex neural networks.
Figure 10. Machine learning workflow: (A) Steps involved in developing AI/ML-based models; (B) Different levels of complexity of ML models, ranging from simple linear regression models to more complex neural networks.
Pharmaceutics 17 01605 g010
Figure 11. Schematic representation of an Artificial neural network (ANN).
Figure 11. Schematic representation of an Artificial neural network (ANN).
Pharmaceutics 17 01605 g011
Figure 12. Regulatory frameworks for the implementation of ML in pharmaceutical spray drying [304].
Figure 12. Regulatory frameworks for the implementation of ML in pharmaceutical spray drying [304].
Pharmaceutics 17 01605 g012
Figure 13. Schematic representation of a hybrid CFD-ML model: (A) CFD-based generated datasets approach; (B) ML model surrogate approach [333].
Figure 13. Schematic representation of a hybrid CFD-ML model: (A) CFD-based generated datasets approach; (B) ML model surrogate approach [333].
Pharmaceutics 17 01605 g013
Figure 14. Digital twin approach for predicting energy efficiency respectively [337], reproduced with permission from Elsevier, 2025.
Figure 14. Digital twin approach for predicting energy efficiency respectively [337], reproduced with permission from Elsevier, 2025.
Pharmaceutics 17 01605 g014
Table 1. Summary of different drying methods and their limitations.
Table 1. Summary of different drying methods and their limitations.
Drying TechnologyProcess DescriptionDrawbacksRefs.
Freeze dryingProducts are frozen and subjected to a vacuum to maintain product qualityHigh processing time (24–48 h)
High production costs and energy consumption
[12,13,14,15]
Fluidized bed dryingParticles are suspended and mixed in a hot air stream for constant and efficient dryingLow product quality
Drying temperature limitations
[16,17,18,19]
Spray–freeze dryingThe product is sprayed, then frozen, and finally dried under vacuumTime-consuming (3 steps)
Complex
Expensive
Fragile particles
[20,21]
Electro sprayingLiquids are sprayed into fine droplets using an electric field, and dried by evaporationReduced production efficiency
High cost
[22,23]
Solar dryingDirect sunlight is used for drying productsClimate dependent
Additional heat and large area requirement
Non-uniform drying
[24,25,26,27]
Superheated-steam dryingDrying of products by heating the steam above its boiling pointUnwanted color changes in products
Temporary moisture increase
[28,29,30,31,32]
Infrared dryingDrying of products takes place using thermal radiationWeak penetrative ability
Product overheating and burning
[33,34,35,36]
Supercritical fluid dryingSupercritical fluids, such as CO2, are used to remove the water from the productHigh cost
Requires organic solvent to improve water solubility and drying efficiency
[37,38,39,40,41,42]
Table 2. Summary of studies applying the DoE approach in pharmaceutical spray drying.
Table 2. Summary of studies applying the DoE approach in pharmaceutical spray drying.
Design MethodologyAPI & ExcipientCritical Process ParametersCritical Quality AttributesRefs.
Response Surface MethodologyHuman type 5 adenoviral vector vaccine, mannitol/dextranInlet temperature, feed flow rate, and feed concentrationResidual moisture content and process yield[160,161]
23 factorial designIvermectin, L-leucine Inlet temperature, feed rate, and atomization air flow ratePowder yield, particle size, and morphology[162,163,164,165]
Box–Behnken designResveratrol, low-methoxyl pectin (LMP), and caprylic/capric glycerides (CCG)Feed rate, inlet temperature, Solute concentration (%w/v), outlet temperature, and solvent concentration (%v/v)Moisture content, particle size, powder yield, and particle size distribution[165,166,167,168,169]
24 full-factorial designDisodium cromoglycate, mannitolFeed rate, feed concentration, inlet temperature, and drying gas flow rateParticle size distribution, powder yield, residual solvent content, and outlet temperature.[170,171]
Half-factorial designBacteriophage MS2 VLP-based candidate vaccine, mannitol, l-leucine, trehalose, and dextranEffect of excipient ratio, feed rate, and atomization pressureParticle size, moisture content, yield[172,173]
Central composite face-centered design (CCFD)Fenofibrate, mannitol, and trehaloseRatio of two carriers, crystallinity of spray-dried powder, and solvent ratioParticle size, batch yield, and antioxidant and antimicrobial activity[174,175]
Circumscribed central composite designDiazepam, mannitolWater/organic solvent ratio, liquid feed flow rate, total solid content, atomizing air flow rate, and type of organic solventDissolution rate, yield, actual drug load, particle size, and crystallinity of drug and excipient[176]
3 × 4 full factorial designCationic liposomal adjuvant formulation 01 (CAF01), lactose, mannitol, and trehaloseChoice of stabilizing excipient and the lipid concentrationYield, moisture content, polydispersity index, particle size, and particle morphology[177]
Factorial 2 × 2 × 3 experimental designEnhanced green fluorescent protein (EGFP) and luciferace (FLuc) Dicer substrate asymmetric duplex siRNAs, trehalose, lactose, and mannitolExcipient concentration and the ratio of nanoparticle to excipientMoisture content, particle morphology, particle size, and powder yield[178]
Table 3. Summary of recent research studies using CFD tools in various spray drying applications.
Table 3. Summary of recent research studies using CFD tools in various spray drying applications.
ProductApplicationResearch AimProcess ParametersMain FindingsRefs.
Tiotropium bromide nanoliposomes (trojan)PharmaceuticalTo produce Tiotropium bromide nanoliposomal dry powder for inhalation using the thin-film hydration and spray drying methodInlet temperature (°C)—110 ± 5
Aspiration capacity (%)—85
Feed rate (mL/min)—10
Newly produced Trojan dry powder showed great promise for the treatment of respiratory diseases
CFD analysis showed that higher inhalation flows increase particle deposition in airways due to greater inertia and turbulence
[199]
Fresh wheyFoodTo fill the gap in the literature between fluid dynamics and spray drying of fresh whey, and provide a detailed assessment of important process parameters and design propertiesInlet temperature (K)—413.15; 453.15; 493.15; 513.15; 573.15 The lowest drying temperature is the optimal process condition, as it resulted in the lowest particle deposition and the highest thermal efficiency
Parameters such as air flow rate and nozzle need to be further explored to understand their influence on thermal efficiency and powder recovery
[200]
Nanostructured silica particleChemicalTo properly evaluate the spray drying process, with a specific focus on nanostructured silica particle formation from sodium silicate precursorInlet temperature (K)—473, 673, and 873
Atomization flow rate (L/min)—2, 4, and 6
CFD simulations properly established the evaporation rate constant (K) and the temperature difference (ΔT) inside the droplets.[201]
Mold powder slurryChemicalTo explain mold powder spray drying using a mathematical model that describes droplet and granule movement as well as heat and mass transfer, and explore the influence of process parameters on granule size using CFD softwareInlet temperature (K)—673–1073
Atomization pressure (MPa)—1.5–2.3
Slurry mass flow rate (kg s−1)—0.01–0.1
Effect of spray drying process parameters such as inlet temperature, atomization pressure, and slurry mass flow rate on the final granule was studied using the CFD model, and results suggested that slurry mass flow rate largely influenced the final granule size[202]
Guava juice formulationFoodTo model the spray-drying of a formulation of guava juice using a CFD approachMass flow rate (kg/h)—81
Inlet temperature (K)—453
Simulation results suggest the need to implement drying kinetics with experimental support, and suitable inlet flow conditions to perform detailed CFD simulations of the spray-drying of fruit juices[203]
SiO2 and ZnO particlesChemicalTo study the modelling of ZnO-SiO2 Composite through a consecutive electrospray and spray drying method Atomization flow rate (L/min)—2–10Applied voltage and the precursor flow rate effectively affected the composite droplet size, whereas gas flow rate and inlet temperature influenced the effectiveness of the composite particle formation in the spray drying process.[204]
Lime slurryFoodTo investigate the applicability of three drying models and three turbulence models in analyzing the drying process in a laboratory spray dry scrubberInlet temperature (°C)—108.8; 130.3; 142.5
Mass flow rate (kg/s)—1.1
Hindered drying mechanistic model is superior to the commonly used d2 law and perfect shrinkage models, and therefore should be preferable in drying applications[205]
L-leucinePharmaceuticalTo develop a new CFD-based model of complex transport and droplet drying kinetics in a lab-scale spray dryer, and relate CFD-predicted drying parameters to powder aerosolization performance from a reference dry powder inhaler (DPI)Liquid feed rate (%)—100
Gas flow rate (L/min)—120
Reducing the CFD-predicted maximum drying rate experienced by droplets improved the aerosolization performance of the powders aerosolized with a reference DPI[206]
Skim milk powderFoodTo simulate the particle movement within the spray dryer, and account for the observed stickiness of the skimmed milk powderAtomization flow rate (kg/s)—0.0139
Inlet temperature (°C)—195
CFD simulations provided accurate predictions of the activity inside the spray dryer, and they agreed with the experimental results
Stickiness arises during the spray drying process, and it can be utilized to control particle agglomeration to obtain better quality powder
[207]
Table 4. Characteristics of ANSYS Fluent and OpenFOAM platforms.
Table 4. Characteristics of ANSYS Fluent and OpenFOAM platforms.
CharacteristicsANSYS FluentOpenFOAMRefs.
Computational costHighLow, but may compromise accuracy[216,217]
AccuracyHighComparable[218,219]
UsabilityUser-friendly, detailed guides and tutorialsLimited, harder learning curve for new users[220,221]
Numerical methodsFinite-Volume-Method (FVM) and electromagnetic equations with Finite-Difference-Method (FDM)Finite-Volume-Method (FVM) for all equations[222,223]
Multiphase modelingGeo-reconstruct scheme, implicit approachDifferent approaches to interface compression, explicit solutions[224,225]
CustomizationUser-defined functionsHighly customizable, enables integration of latest numerical methods[217,220]
Computational resourcesHigh computational resources and knowledge requiredLess resource-intensive[217,226]
Table 6. Comparison of prediction accuracies of different ML models in the spray drying process.
Table 6. Comparison of prediction accuracies of different ML models in the spray drying process.
Research aimML AlgorithmDatasetPrediction AccuracyRefs.
Physical stability of solid dispersions at 3 months and 6 monthsRandom forest (RF), LightGBM,
Support vector regression (SVR)
Fifty drug compounds with ten molecular descriptors RF—82.5% (highest)
LightGBM—80.83%
SVM—77.50%
[288]
Prediction of median spray-dried dispersion particle size (SDD)Partial least square (PLS), Support vector regression, Neural networks (multi-layer perceptron)680 SDD lots using 57 different pressure nozzles across two spray dryer scales and 88 unique APIsPLS—7.69 and 6.81 (training and testing RMSE)
SVR—5.39 and 6.56
Neural network—3.72 and 6.10
[131]
To obtain an effective solid dispersion formulation design for the oral administration of water-insoluble drugsRF, SVR, LightGBMThree data sets–physical stability (646 lots), dissolution curves (702 lots), dissolution profiles (4214 samples)RF—77.7%
LightGBM—76.4%
SVR—59.7%
[289]
Prediction and fault detection in key performance parameters for a milk spray drying process plantDecision tree, random forest, logistic regression, and SVM17400-time instancesDecision tree—99.85%
Logistic regression—99.59%
Random forest—99.85%
SVM—95.40%
[290]
Table 7. Summary of the combination of DoE and ML models used in the spray drying process.
Table 7. Summary of the combination of DoE and ML models used in the spray drying process.
DoEML ModelProductProcess ParametersResponse VariablesPrediction AccuracyRefs.
Response surface methodology (RSM)Artificial Neural Network (ANN)Aripiprazole -cyclodextrin complexFeed rate, inlet temperature, feed concentration, compressed air flow rate, and aspirator capacityPowder yield, moisture contentR2 = 0.854 for yield, R2 = 0.886 for moisture content[293]
24 full-factorial designExtreme gradient boostingα-lactose monohydrate Feed rate, outlet temperature (Tout), and solid concentration.Powder yield, residual moisture content, cut off diameter (X50)R2 = 0.982 for yield,
R2 = 0.998 for residual moisture content, R2 = 0.923 for X50
[294]
Quality by design (QbD)Random ForestDeionized waterAspirator rate, inlet temperatureOutlet temperature (Tout)R2 = 0.99[295]
DoE–4 factors (two continuous and two categorical factors)Support vector machine (SVM) and ANNLactose/Polyvinylpyrrolidone and lactose/Kollidon physical mixturesType of core and shell materials and their concentrationsPowder compactibilityRoot mean square error (RMSE) = 2.3% for ANN,
RMSE = 6.8% for SVM
[292]
Table 8. Key elements of TensorFlow and MATLAB machine learning platforms.
Table 8. Key elements of TensorFlow and MATLAB machine learning platforms.
FeaturesTensorFlowMATLABRefs.
ApplicationTrains advanced neural networks for accurate predictions and process optimization in spray drying process Mathematical modeling, simulation of drying kinetics, and experimental data analysis[217,298]
FlexibilityHighly flexibleLess flexible and inefficient in handling advanced neural networks[296,297]
ScalabilityHighly scalable for large and complex datasetsLimited due to model dimensionality[296,297]
Experimental analysisNot generally used for direct experimental validationEffective for image processing and data analysis[299,300]
Computational efficiencySuitable for large-scale complex computationsLess suitable for complex system modeling[297,299]
Table 9. Key FDA/EMA aspects in pharmaceutical drug manufacturing.
Table 9. Key FDA/EMA aspects in pharmaceutical drug manufacturing.
ElementFDAEMARefs.
AI/ML in drug manufacturingDiscussion paper, lifecycle management, SaMD frameworkAI work plan till 2028, valid AI algorithms for data analysis in Pharmacopeia[309]
Algorithm categoriesLocked vs. adaptive/continuous learning Endorsement of AI for data evaluation[309]
Main emphasis areasBias mitigation, transparency, real-world monitoringResponsible utilization, risk management, training[310]
International CollaborationJoint good practice
principles with Health Canada and the UK’s MHRA
Partnership with HMA (Head of Medicines Agencies), open access to methodologies[311,312]
Regulatory statusProgressing, with explicit guidance expectedOngoing, with published work plans and guidelines[313]
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

Wahab, W.; Alshamsi, R.; Alharsousi, B.; Alnuaimi, M.; Alhammadi, Z.; Al-Zaitone, B. Recent Developments in Pharmaceutical Spray Drying: Modeling, Process Optimization, and Emerging Trends with Machine Learning. Pharmaceutics 2025, 17, 1605. https://doi.org/10.3390/pharmaceutics17121605

AMA Style

Wahab W, Alshamsi R, Alharsousi B, Alnuaimi M, Alhammadi Z, Al-Zaitone B. Recent Developments in Pharmaceutical Spray Drying: Modeling, Process Optimization, and Emerging Trends with Machine Learning. Pharmaceutics. 2025; 17(12):1605. https://doi.org/10.3390/pharmaceutics17121605

Chicago/Turabian Style

Wahab, Waasif, Raya Alshamsi, Bouta Alharsousi, Manar Alnuaimi, Zaina Alhammadi, and Belal Al-Zaitone. 2025. "Recent Developments in Pharmaceutical Spray Drying: Modeling, Process Optimization, and Emerging Trends with Machine Learning" Pharmaceutics 17, no. 12: 1605. https://doi.org/10.3390/pharmaceutics17121605

APA Style

Wahab, W., Alshamsi, R., Alharsousi, B., Alnuaimi, M., Alhammadi, Z., & Al-Zaitone, B. (2025). Recent Developments in Pharmaceutical Spray Drying: Modeling, Process Optimization, and Emerging Trends with Machine Learning. Pharmaceutics, 17(12), 1605. https://doi.org/10.3390/pharmaceutics17121605

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