Journal of Food Engineering: Kosana Pravallika, Snehasis Chakraborty, Rekha S. Singhal

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Food Engineering 343 (2023) 111375

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Review

Supercritical drying of food products: An insightful review


Kosana Pravallika, Snehasis Chakraborty *, Rekha S. Singhal
Department of Food Engineering and Technology, Institute of Chemical Technology, Matunga, Mumbai - 400 019, India

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

Keywords: Supercritical drying (SCD) has emerged as an alternative technology to conventional drying technologies. SCD
Supercritical drying addresses the challenges faced by food markets by providing high nutrient retention and superior quality
Supercritical fluids products that resemble original products. Supercritical carbon dioxide (SCCO2) is a widely used solvent that
Microbial inactivation
solubilizes water in the sample and gets separated during decompression. Cosolvent (ethanol) is used prior to the
Enzyme stability
Quality attributes
application of supercritical fluid to increase the drying efficiency of the SCD process. Examples include spices, a
Shelf life few fruits and vegetables, and other bioactive edibles/gels. This paper provides a detailed discussion on the SCD
operational assembly, the mechanism of mass transfer phenomena, and factors affecting SCD. The effect of SCD
on various physicochemical and sensory attributes, rehydration capacity, and shelf life of SC dried samples.
Finally, the challenges and future scope of SCD have been summarized. This review delivers comprehensive and
updated information on advancements in SCD for food products.

1. Introduction forces, which may disrupt the microstructure (Lee et al., 2011; Vetralla
et al., 2018). Microwave drying enhances the drying rate but causes
Drying is one of the oldest preservation technologies employed in excessive dehydration of the product due to improper heat and mass
food processing and preservation of many predominant agricultural transfer control. Other thermal drying techniques, such as spray drying,
products such as cereals, pulses, fruits, and vegetables. It extends the employ high temperatures (approx100–210 ◦ C) during processing
shelf life by lowering the water activity and thus inhibiting the growth of (Suravanichnirachorn et al., 2018). The quality of spray-dried powders
microorganisms. Inadequate drying of the product may decrease shelf is highly dependent on spray drying operating parameters such as inlet
life due to quality degradation and pose food safety risks. Besides the temperature, feed flow rate, atomizer speed, type of carrier, and carrier
shelf life extension, drying should retain the original texture, color, concentration. It might lead to the loss of heat-sensitive compounds (e.
flavor, and nutritional content of the food product (Djekic et al., 2018). g., emulsions and fruit powders) (Santhalakshmy et al., 2015). To
Furthermore, drying reduces the product’s weight and volume, which overcome the heat damage, an alternative is to remove water by
aids in better storage and transportation (Lee et al., 2011). Consumer freeze-drying in which the water is frozen, followed by sublimation of
awareness of the health benefits of food products has propelled a de­ the ice to a vapor phase by manipulating pressure and temperature. It is
mand for products with high nutrient retention and better organoleptic not commercially economical because of its high cost of operation, low
properties (Batiha et al., 2021). However, there are several limitations to production capacity and yield. Vacuum drying is similar to that of
employing conventional drying when the product is rich in thermo­ freeze-drying which involves the whole process except the freezing stage
sensitive compounds (Lee et al., 2011). Sun drying has limitations of (e.g., fruits, vegetables, high-quality juice powders). Vacuum drying
longer drying time and non-uniform drying rate. Industrial dryers are could disrupt the structure of the food product (Vetralla et al., 2018).
primarily convective, where hot air or gases are used as the heat transfer Supercritical drying (SCD) is one of the emerging technologies that
medium (Moses et al., 2014). Contact drying (around 105–150 ◦ C) and can remove water from the food at a state above the critical conditions of
air drying (typically 65–85 ◦ C) are mainly preferred due to their low cost the solvent used without disturbing the structure of the product. SCD
and ease of operation. Due to the application of high temperatures, these develops a product with high physicochemical stability similar to its
may damage color, texture, and other sensory properties. It involves the original nature with better microbial, enzymatic, and shelf stability
phase transition and mass movement of water that develops capillary (Moses et al., 2014). The process time of SCD is very less due to the

* Corresponding author. Food Engineering and Technology Department, Institute of Chemical Technology, Matunga (E), Mumbai, India - 400 019.
E-mail addresses: fet21k.pravallika@pg.ictmumbai.edu.in (K. Pravallika), sc.chakraborty@ictmumbai.edu.in, snehasisftbe@gmail.com (S. Chakraborty),
rsinghal7@rediffmail.com, rs.singhal@ictmumbai.edu.in (R.S. Singhal).

https://doi.org/10.1016/j.jfoodeng.2022.111375
Received 27 August 2022; Received in revised form 6 November 2022; Accepted 6 December 2022
Available online 7 December 2022
0260-8774/© 2022 Elsevier Ltd. All rights reserved.
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

involvement of both temperature and pressure, and the cost of operation 2018). However, a complete and consolidated review of the operation of
is also lower compared to advanced drying technologies such as spray SCD, the mechanism of moisture removal, and its effect on various
drying, vacuum drying, and freeze-drying. Hence it presents many ad­ quality attributes and shelf life of food products have not been under­
vantages over conventional drying techniques (Lopes et al., 2018). In the taken till date. Hence, the objective of the review is to collate the
SCD process, the solvents used are termed as supercritical fluids (SCF) updated literature on the process and operational assembly of SCD,
that solubilize the water in the sample and further SCF-water mixture factors affecting SCD, the effect of SCD on microbial and enzymatic
gets separated due to decompression (Benali et al., 2014). In the early inactivation in various solid foods, and the effect of SCD on various at­
1980s, supercritical fluids have been employed primarily for the tributes (physicochemical, rehydration capacity, bioactive, sensory at­
extraction process in the food industry. The process of supercritical fluid tributes), and the shelf life of the product.
extraction (SCFE) and supercritical drying (SCD) is the same, but the
results are different. For example, the former process involves extracting 2. Materials and methods
desired components from the sample (such as bioactive from fruit
pomace or oil from seed cake), whereas the latter involves removing Due to the fact that supercritical drying technology is still in its in­
water from the sample that gives a product similar to other drying fancy, there aren’t many articles on it. The first step was to search the
techniques. High-quality fruits and vegetables, spices, and meat prod­ articles using the keyword “supercritical drying of foods.” Roughly 25
ucts have been generally preferred to be dried using SCD. papers were found. Writing a review article based on very few publi­
In the last couple of decades, researchers have been exploring the cations and drawing conclusions is exceedingly challenging. Hence the
applications of SCD on various food products such as apples (Djekic keyword was modified to “supercritical drying and supercritical,” and
et al., 2018; Lee et al., 2011; Vetralla et al., 2018; Zambon et al., 2021), all the publications that were then available were either related to su­
carrots (Brown et al., 2008), coriander (Bourdoux et al., 2018; Michelino percritical fluid extraction or had a drying background in pharmaceu­
et al., 2018; Zambon et al., 2018), chicken breast (Morbiato et al., 2019), tics. This paper covers research and review articles from the last 15
basil (Bušić et al., 2014), and red bell pepper (Zambon et al., 2020). A years, from 2007 to 2022.
few reviews are available on a brief description of SCD (Benali et al.,
2014; Bourdoux et al., 2018) and the application of SCFs as a drying aid 3. Supercritical drying as a technology
and its effect on a few foods ( Nautiyal, 2016). The detailed effect of SCF
pasteurization on the mechanism of microbial and enzymatic inactiva­ 3.1. Process of supercritical drying and its assembly
tion in various foods has been summarized in few literatures (Spi­
limbergo et al., 2018; Ferrentino and Spilimbergo, 2011; Zambon et al., Supercritical drying is the process of removal of water from the

Fig. 1. Schematic showing (a) sequential steps of supercritical drying (SCD) (Zambon et al., 2021): (b) mechanism of water removal from food matrix by SCD
(Chakraborty et al., 2020): (c) comparison of SCF and cosolvent + SCF in moisture removal during SCD (Vetralla et al., 2018).

2
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

product using SCF and cosolvent by manipulating temperature and pressure is completely reduced to 1 atm (Wang et al., 2021 Zheng et al.,
pressure for predetermined time. As depicted in Fig. 1a, SCD occurs in 2010; Brown et al., 2008; Vetralla et al., 2018).
following steps: a) solubilization of cosolvent into the sample, b) pres­
surizing the cosolvent mixed sample with supercritical solvent, and c) 3.2. Solvents and cosolvents used in supercritical drying
then finally depressurizing to remove water along with cosolvent and
solvent. A set of solid foods such as carrots, apples, coriander, red bell Supercritical fluids have piqued the interest of researchers due to
pepper, and chicken breast have been dried using SCD. their exceptional properties such as high diffusion rate (10 times greater
As shown in Fig. 2, the drying assembly of SCD is different from the than gas), low viscosity (10–100 times lower than liquid), low surface
conventional drying process. Typically, an SCD assembly consists of a tension, higher density than gas and liquids, making SCF a good solvent
drying chamber, solvent feed cylinder, desiccator tube, heat exchanger (Girotra et al., 2013)). SCF can permeate through the cell membranes
(condenser, cooler, and heater), pressure gauge, test flow meter, and and penetrate deep into the sample, and dissolve water. Many solvents
cosolvent feed cylinder (HPLC). The food sample such as carrot (90% such as carbon dioxide (CO2), nitrogen, ammonia, and freon have been
moistened, cylindrical pieces of 25 mm height and 4 mm diameter) is used as supercritical solvents. Among all the SCFs, CO2 is highly
positioned in a stainless-steel mesh sample holder and placed at the preferred in food applications over other solvents since other solvents
centre of the drying vessel, which is well equipped with a temperature and cosolvents have a negative impact on consumer health and on the
and pressure controlling device. An HPLC pump supplies the precise environment (Table 1) (Brunner, 2005). The supercritical nature of CO2
amount of cosolvent (6% ethanol) (Brown et al., 2008). It requires an is depicted in Fig. 3. CO2 is specified as a generally recognized as safe
extra processing time of 30 min for the complete dissolution of water in (GRAS) solvent by the USFDA and fulfils all the ideal properties of SCF
cosolvent. After 30 min, the solvent (mostly CO2) is supplied from the (Corzzini et al., 2017). With low critical parameters (31.25 ◦ C/7.38
cylinder through a desiccator tube and cooled to a liquid state for proper MPa), the generation of SCCO2 becomes easier and makes it suitable for
flow, and then compressed to the desired pressure (20 MPa). The pres­ drying without damaging the structure of the sample (Harrison, 2020;
surized solvent is preheated to the desired temperature (40–60 ◦ C) with Nowacka et al., 2021). Additionally, it is naturally abundant, inexpen­
the help of a heat exchanger. Once the necessary temperature and sive, non-flammable, and can also be recycled easily without causing
pressure are established, the micrometering valve opens, and subse­ global warming (Zheng et al., 2010). CO2 has a dipole moment of zero;
quently, SCF is fed to the drying chamber with a constant flow rate of hence it is considered as a non-polar solvent and limits the solubility of
~2 L/min at STP. A backpressure regulator valve maintains constant polar compounds, but the slight polarity is because of the quadrupole
pressure throughout the process. Due to the pressure exerted by the SCF moment that allows the CO2 to dissolve a few polar compounds at high
flushing, the cosolvent-water mixture oozes out of the sample, owing to pressure.
its high solubility in a short contact time. The flow rate is continuously An additional agent called cosolvent is used before SCF to improve
controlled and monitored by a wet test flow meter throughout the the drying efficiency by increasing the solubility of polar solutes and
process. The SCF gets saturated due to the constant cosolvent-water decreasing the processing time (Nowacka et al., 2021). Organic solvents
mixture pushing out of solid food matrices. After 10 min of SCF flush­ such as alcohols (ethanol and methanol), ketones (acetone), and
ing, the micrometering escape valve opens, and SCF moves out at a aliphatic alkanes are employed as supercritical cosolvents in the SCD
depressurization rate of 0.4 MPa/min by maintaining a constant tem­ process (Table 1) (Zheng et al., 2010). Among these, ethanol is
perature (typically isothermal depressurization). It expands and leads to GRAS-compliant and is an excellent cosolvent for usage in the food
two phases: solvent-rich vapor phase and cosolvent-rich liquid phase. sector. Ethanol is highly preferred for making SCF to dissolve polar
The solvent-rich phase is collected and cooled below its critical tem­ compounds such as water due to the hydrogen and dipole interactions
perature to convert to liquid using a heat exchanger, allowing it to use between ethanol and SCF (Vatai et al., 2009). Acetone is highly flam­
for a few more cycles. The process is performed till the cosolvent in the mable because of its low flash point of – 20 ◦ C, due to which a small
pores of the sample gets removed completely (Benali et al., 2014). The quantity in the air leads to explosions and fire accidents. Methanol is
water in cosolvent is removed by passing it through solid water adsor­ considered toxic, especially in the food processing sector (Şahin et al.,
bents or by bubbling through liquid desiccants (Zambon et al., 2021). 2018).
Finally, the dried sample is collected from the drying chamber after the

Fig. 2. Schematic depiction of SCD apparatus (Brown et al., 2008).

3
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Table 1
Characteristic properties of some important solvents and cosolvents used in supercritical drying of food products.
Supercritical Critical temperature Critical pressure Characteristic property Limitations References
fluid (◦ C) (MPa)

Carbon dioxide 31.25 7.38 ▪Non-toxic ▪The low solubility of polar compounds Harrison (2020)
▪Does not impart taste or color
▪Recyclable
▪Easily accessible
▪Low critical temperature and pressure
Nitrogen − 147 3.35 ▪Moisture carrier at very high ▪Very low critical temperature Benali et al.
temperatures (2014)
Ammonia 132 11.28 ▪Good solvent because of low critical ▪Caustic and dangerous ▪ Distinct odor Zheng et al.
conditions (pungent smell) (2010)
▪High polarity
Nitric oxide 36.15 7.285 ▪The dipolar moment is similar to CO2 ▪Has an anesthetic effect Benali et al.
▪Causes oxidation (2014)
Freon 27 3.5–4.0 ▪Low critical temperature and pressure ▪Highly explosive Khaw et al.
▪Easy accessibility ▪Causes global warming (2017)
▪Not used in food applications ▪Causes organ damage
Water 374.15 22.06 ▪High polarity ▪High critical temperature Benali et al.
▪Accessible ▪Powerful oxidizer (2014)
Ethanol 243 6.3 ▪Solubilizes both polar and non-polar ▪High critical temperature Harrison (2020)
substances ▪Persistence of residue in the final product
▪Highly explosive

Fig. 3. Pressure and temperature relationship of a pure substance (Brown et al., 2008). Tc, Pc: critical temperature and pressure and Ttp, Ptp: triple point temperature
and triple point pressure.

3.3. Overview of mass transfer phenomena during SCD pore is lower than the mean free path of the cosolvent/solvent mole­
cules. Hence, molecules during diffusion collide with the walls of the
Subjecting the food samples to SCD involves mass transfer opera­ pores rather than the other molecules, which is well significant in the
tions. Water can be found in the food sample as free water (squeezed out micropores. The cosolvent expands due to the dissolution of SCF into
with external force), bound water (tightly bound with other components cosolvent, which causes syneresis of the cosolvent along with water from
such as protein, fat, vitamins, pectin, etc.), and entrapped water (typi­ the sample. As the drying continues, the concentration of cosolvent
cally present as free water in the innermost layer such as in a matrix and between the sample and SCF stream decreases with time, and eventually
is removed as free water on cutting/squeezing) (Lewicki, 2004). Free leads to the complete replacement of all the liquid by SCF. The flow
and entrapped water can be removed easily compared to bound water. direction of solvent and cosolvent is also very important. Downflow,
When flushed with a jet of cosolvent at a supercritical state, it primarily up-flow, and horizontal flow are typically three flow directions. Up-flow
dissolves the free water present in the sample, then moves into deeper and horizontal flow take a longer drying time as it is very difficult to
layers and dissolves the entrapped water, and then the bound water. push the cosolvent by the low-density solvent. Hence, downflow is more
Water molecules detach from the active sites of the cell components and effective due to pushing out of cosolvent immediately by passing
diffuse out through multiple layers of the food sample (Chakraborty high-pressure SCF without a back mixing (Derossi et al., 2008). It is
et al., 2020). When the stream of SCF is flushed onto the sample, the extremely difficult to develop kinetic modeling based on these few
concentration gradient of cosolvent between the surface of the sample studies conducted on wide variability of food categories.
and the SCF stream becomes high. Hence, there is a very high convective
mass flux of the cosolvent from the sample into the SCF stream. SCF 3.4. Selective dissolution of water by SCF and cosolvent
diffuses deep towards the core of the sample (Fig. 1b). The diffusion
mechanism of pores follows Knudsen and molecular diffusion. The pores The food sample is a complex mixture of different nutrients and
inside the food samples are in the range of mesopores or micropores. bioactives. Hence, the removal of water-soluble nutrients and bioactives
Molecular diffusion governs the diffusion in mesopores as the size of the along with water with the application of SCF is undesirable. Hence
pore is larger than the mean free path of cosolvent/solvent molecules. before its application, it is important to know the reaction between food
Knudsen diffusion governs the diffusion in micropores as the size of the with SCF. The solubility of water is predicted to be 1.8–2.45 mg/g at

4
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

40 ◦ C/10 MPa and the solubility of CO2 in water is 5.1 g CO2/100 g pressure increases the solubility of water in compressed SCCO2. When
water at 37 ◦ C/10 MPa (Wrona et al., 2017). The extractability of any the carrot slices were subjected to 40 ◦ C/20 MPa/90 min, the moisture
compound depends on the presence of specific groups in the structure, as content was reduced to 50% of the initial moisture content. When basil
well as their polarity and molecular weight. Previous studies have leaves were dried using CO2 at 40 ◦ C/8–10 MPa/2 h, it gave a product
revealed that low molecular weight (<250 MW) organic compounds with 5.4–5.7% moisture content (Bušić et al., 2014).
such as esters, aldehydes, ethers, ketones, lactones, and epoxides exhibit
excellent solubility in SCCO2 and are favored by low pressures (7.5–10 3.5.3. Solvent flow rate
MPa). Moderately polar compounds with a molecular weight of The flow rate of the solvent regulates the diffusivity flux of solute
250–400, such as terpenes, sesquiterpenes, and other benzene derivates, transfer into the solvent in a supercritical state. The increased velocity of
are moderately soluble in SCCO2 under high-pressure conditions. The the solvent helps in deeper penetration of the solvent and induces mass
polar compounds of molecular weight ≥400 that include carboxylic acid transfer by forced convection (Nowacka et al., 2021). However, the
and hydroxyl groups such as sugars, tannins, proteins, carotenoids, an­ residence time is negatively affected, and the term dilution factor (flow
thocyanins, tocopherols, B-complex vitamins, and waxes are barely rate to the solvent volume) comes into discussion. The dilution rate
soluble in SCCO2 (Şahin et al., 2018). The solubility of high molecular should be minimum to dissolve the water from the sample. Hence, the
weight compounds decreases with an increase in the pressure of the flow rate should be optimized in a certain way for effective moisture
process. The diffusion rate of the high molecular weight compound from removal. It should be controlled for a better drying process. For instance,
the internal portion of the sample is lower due to the resistance offered the drying time was reduced from 300 min to 210 min by maintaining a
by pores (Uwineza and Waśkiewicz, 2020). flow rate of 2 L/min with 6% ethanol as a cosolvent in carrots (Brown
Additionally, mass transfer resistance created by the sample’s et al., 2008).
permeability of the cell wall and membrane also affects the water
extractability. The cell wall is permeable, while the cell membrane is 3.5.4. Cosolvent
highly selectively permeable to polar compounds, which controls the Cosolvent plays a crucial role in increasing the drying efficiency of
movement of solute molecules (Phisut, 2012). Since CO2 and ethanol SCD. The rate of removal of moisture from the sample increases with an
can penetrate through the cell wall and membrane because of their increase in the concentration of ethanol up to the optimum range. As
non-polar nature, the polar solutes present in the sample cannot move shown in Fig. 1c, when the carrot slices were subjected to SCD for 90
out of the cell membrane. Even though water is a polar solute, the min, the moisture removal by SCCO2 was 50% of the initial moisture of
membrane allows water from high potential (internal portion) to low the sample. While the samples treated with ethanol did not show any
potential (outside the sample) to equalize the gradient. The processing effect on the sample’s moisture content for the first 30 min because the
conditions such as temperature, pressure, flow rate, and other parame­ ethanol mixed with the water and settled in the pores of the sample. The
ters modify the permeability of the membrane that limits the diffusion of ethanol and water mixture spills out of carrot pores by the pressure
other polar solutes (Phisut, 2012; Thorat et al., 2012). These few liter­ created by SCCO2. Further, the moisture content became much lower
ature reports suggest that the loss of polar solutes from the sample along than in the SCCO2-treated sample (Brown et al., 2008). Vetralla et al.
with the water during drying can be controlled. (2018) studied the effects of SCD conducted at 35 ◦ C/0.001 MPa on
apple slices using SCCO2 and ethanol-SCCO2. They concluded that the
3.5. Factors affecting the efficacy of supercritical drying pre-treatment of the sample with ethanol could further decrease the
drying time of SCD. Treatment of the sample solely with SCCO2 for 120
The efficacy of the SCD process depends upon the process tempera­ min decreased moisture to 18.5 ± 1.5%, while the pre-treatment with
ture and pressure, which affect the density of the solvent, type and ethanol for 40 min followed by flushing with SCCO2 for 10 min reduced
concentration of solvent and cosolvent, and the size of the sample. Since it to 11.6 ± 2.5%. No further decrease in moisture content was observed
research has been carried out on various categories of foods such as when the pre-treatment time of ethanol and SCCO2 was increased from
carrots, apples, red bell pepper, and basil leaves, among many others, 50 to 80 min and 10–40 min, respectively. The shrinkage was more in
which are different in moisture content, morphological structure, ethanol-SCCO2, and it also had a broader pore size than SCCO2 dried and
texture, and composition, it is difficult to come up with a general control apple slices. This is because of the involvement of specific
conclusion. But these are the major factors, and their effect is significant physical and chemical interactions between the cosolvent and the water
in foods. Hence, these have to be optimized/controlled while designing due to strong hydrogen bonding and dipole-dipole interactions.
the treatment condition for better results. Furthermore, the SCCO2 could reduce moisture content by 80%, indi­
cating that SCD is suitable for drying apples. However, the addition of
3.5.1. Process temperature cosolvent increases the drying rate and reduces the dehydration time
The process temperature is one of the most critical factors that in­ (Lee et al., 2011).
fluence the SCD process. The viscosity and surface tension of the SCF
decreases on increasing the temperature of the process, which facilitates 3.5.5. Sample size
the penetration of SCF into the sample (Asafu-Adjaye et al., 2021). The sample size is also an essential factor influencing the drying rate
Furthermore, the diffusivity and solubilizing power of SCF increase at an and drying time. Lee et al. (2016) conducted a study on the drying ki­
elevated temperature (Nowacka et al., 2021; Wason et al., 2021). Brown netics of low and high sample sizes of wet okara. Surprisingly, high
et al. (2008) conducted supercritical drying on carrot slices. They found sample size material (>1000 g) was found to have a linear relationship
that an increase in the process temperature of the SCD (40–60 ◦ C) between drying time and the rate of moisture removal, which indicated
elevated the drying rate. The SCF penetrated deep into the sample due to a constant drying rate period. In the case of a lower sample size (<500
decreased viscosity and enabled the solubilization of more water in SCF. g), the moisture remained in the sample (approached 30%) and the
drying rate decreased with time, which indicated a falling rate drying
3.5.2. Pressure period. Lee et al. (2016) treated the wet okara samples of different
Pressure is also a critical factor in influencing the physical properties sample sizes at 40 ◦ C/10.3 MPa for 10–40 min. They found the water
of SCF. A minimal change in the pressure has drastic modifications to the uptake efficiency of SCCO2 was higher (36.9%) in a larger sample size of
product in a few cases. The higher the pressure created by the SCF, the okara (>1000 g) (0.625 mg water/g CO2) than in a smaller sample size
more is the oozing out of the water and cosolvent from the sample. In (18.6%) of okara (<500 g) (0.315 mg water/g CO2). The passage of only
this case, a change in pressure is more effective than a change in tem­ a fraction of SCCO2 through the matrix of the low sample size wet okara
perature (Lee et al., 2011). Brown et al. (2008) found that an increase in could be the possible reason.

5
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

4. Impact of SCD on various attributes of food products samples, but after 40 min, the rehydration rate was higher in
ethanol-treated SCCO2 dried carrot pieces (Brown et al., 2008). Brown
In the last decade, the effect of SCD has been evaluated on the et al. (2008) also found that samples dried using ethanol-treated SCD
product characteristics of a few food products. These include the impact showed fewer dense structures and better rehydration capacity. The
of SCD on physicochemical properties, microbial stability, enzyme sta­ processing pressure and temperature of SCD did not significantly affect
bility, organoleptic properties, and shelf life of the product. the rehydration capacity. The rehydration process of SC dried samples is
irreversible, as the moisture content of control samples is lower than the
4.1. Physicochemical properties and bioactive compounds moisture content of the rehydrated sample. The reason might be the
structural rigidity of SC-dried samples due to the removal of only water
The physicochemical properties such as moisture content, shrinkage, without disturbing the pore structure and might be due to the larger
micropore size, structure, density, non-enzymatic browning (NEB), pores of SC-dried samples. The rehydration capacity is reported to show
bioactive compounds (carotenoids, ascorbic acid, total phenolic content, hysteresis due to the rupture of cells and structure disruption that occurs
and total flavonoid content) are fundamental and crucial in meeting during drying. The evidence from these few studies indicates that SCD
consumer demands and maintaining the product’s shelf life. The results in products with superior rehydration capacity compared to
changes in these properties after SCD are product specific. Very limited other drying technologies.
information is available on its application in food products, which is Data corresponding to changes in the moisture content of dried
summarized in Table 2. Brown et al. (2008) found that 50 ◦ C/20 MPa/2 samples during rehydration have been quantified in terms of rehydra­
L/min followed by depressurizing to 0.4 MPa/min reduced to half the tion ratio (RR), which is the ratio of the weight of the fresh sample to the
moisture in carrots in 90 min. SCD of carrots enabled the retention of weight of the dried sample. Three semi-empirical models are commonly
their original structure and shape far better than air-dried carrots which used to describe the rehydration kinetics, viz., exponential, Peleg model,
had undergone shrinkage. Samples dried using modified SCD showed and Weibull model, which are summarized in Table 4. These are
fewer dense structures and better rehydration than air-dried ones. They developed to know the effect of the porous microstructure of the dried
also studied agar gels (2% agar w/v) dried using SCCO2 at 50 ◦ C/20 sample on the quality of rehydration (Marques et al., 2009). These
MPa/180 min, reducing moisture to 0.1%. Treatment of red bell pepper models have various parameters which have been analyzed by using
at 60 ◦ C/12 MPa/100 kg/h showed no significant changes in texture or non-linear regression analysis.
shape compared to freeze drying. Similar results were obtained at The exponential model is given by a semi-empirical equation (Eq.
40 ◦ C/14 MPa/160 kg/h/16 h in red bell pepper (Zambon et al., 2020). (1)):
In one report, aerogels were produced from 5, 6 & 7% barley β-glucans ( )
RR = RRe (RRe − 1)exp − K3 tk4 (1)
and dried with different techniques such as freeze drying, air drying, and
ethanol-SCCO2 drying (Comin et al., 2012). The SCD sample had a lower
Whereas RR is the rehydration ratio, RRe is the saturation rehydration
density than air-dried samples and a more consistent structure than
ratio, K3 is the kinetic constant (min− 1), and K4 is a dimensionless
freeze-dried samples. There was no significant effect of the concentra­
constant. The value K4 = 1, and the exponential model leads to first-
tion of β-glucans on density and surface area.
order kinetic expression (Eq. (2)).
Braeuer et al. (2017) conducted a study on the SCD of mango and
persimmon at 40 ◦ C/10 MPa/300 min reduced the water activity to RR = RRe (RRe − 1)exp(− K3 t) (2)
0.026 and 0.124, respectively. They found slight variations in the
This model explains the water absorption during the rehydration
carotenoid content, which could be due to the dissolution of non-polar
process in various fruits and vegetables (Krokida and Marinos-Kouris,
carotenoids in SCCO2. Bušić et al. (2014) conducted a comparative
2003).
study on freeze drying, air drying, and SCD (40 ◦ C/10 MPa/1000 L/h/3
The semi-empirical model proposed by Peleg is described by 2 pa­
h) on the antioxidant and bioactive properties of basil leaves. They
rameters in a non-exponential equation (Eq. (3)). Peleg model estimates
concluded that SCD was on par with freeze drying in retaining chloro­
the moisture content of the sample on a dry basis of the sample as:
phyll, total phenols, flavonoids, and antioxidants in basil leaves. The
betacyanins, betaxanthins, and polyphenols in beetroot are reported to 1 t
RR = RRe − + (3)
be degraded by 58%, 32%, and 30%, respectively, after SCD at 55 ◦ C/60 K2 K1 + K2 t
MPa/30 min (Marszałek et al., 2017). There is not much evidence on the
effect of SCD on physicochemical properties, although extensive in­ Where t indicates the time in minutes, K1 is the Peleg kinetic constant
vestigations are required in this area. with a dimension of time, and K2 is the Peleg capacity constant, which is
dimensionless. It is related to the equilibrium moisture content Xeq as
4.2. Rehydration capacity given in Eq. (4) as follows (Inyang et al., 2018):
1
Rehydration capacity is the essential parameter of dried products, as Xeq = Xo + (4)
K2
they are to be rehydrated for consumption. During SCD, water gets
removed and thus reduces the possibility of engaging the sample with Peleg’s equation has been found to adequately describe the rehy­
water molecules through hydrogen bonding. It helps to form strong dration kinetics of morel and potato cubes (García-Pascual et al., 2006;
sample molecule-molecule interactions and intermolecular polymeric SalimiHizaji et al., 2010).
junctions, resulting in a strong solid microstructure and superior rehy­ The Weibull model has been employed to greatly predict the rehy­
dration capacity (Plazzotta et al., 2019). A few studies on rehydration dration ratio of food systems that have some variability, such as the
profiles of supercritical dried (SCD) products are discussed in Table 3. In varied rehydration rate with respect to time in freeze-dried tomatoes
all the previous studies, it was observed that the uptake of water by the (Lopez-Quiroga et al., 2019). It can be expressed by two parameters: α,
samples was greatly higher during the initial period, but gradually scale parameter (dimensionless), which is the inverse of the rate, and β,
slowed down (Morbiato et al., 2019; Brown et al., 2008). The rapid rise shape factor with a dimension of time (Eq. (5)) (Inyang et al., 2018).
in the initial rehydration rate was due to the filling of intercellular [ ( )β ]
t
spaces and polymeric junctions. As the absorption proceeds, the rehy­ RR = RRe (1 − RRe )exp − (5)
α
dration rate declines. Pure SCCO2 and EtOH–SCCO2 treated carrot pieces
were rehydrated at 50 ◦ C for 5 min didn’t show any difference in The selection of the model for validating experimental data depends
moisture content initially as the rehydration rate was rapid in both the upon the statistical criteria such as coefficient of determination (R2),

6
K. Pravallika et al.
Table 2
Influence of supercritical drying on the physiochemical properties and bioactive compounds of food products.
Food Matrix Treatment conditions Max changes in physiochemical properties Max changes in sensitive bioactive compounds Other key observations Reference

Strawberry 40 ◦ C, 13.3 MPa for 420 min at >92% weight loss aW was <0.34 •100% retention of vitamin C •SCD can be used as sustainable green drying Zambon et al.
19 kg/h •95% retention of total anthocyanin content technology (2022b)
•76% retention of TFC
Strawberry 40 ◦ C, 10 MPa for 360 min 98% moisture removal Zambon et al.
slices 90% weight loss (2022a)
Red bell 40–50 C, 10–14 MPa, with

Reduced the water activity to 0.262 at 50 ◦ C, 12 MPa •Better retention of texture, shape, and rehydration Zambon et al.
pepper 80–220 kg/h for 6/16 h with 100 kg/h at 6 h and at 40 ◦ C, 10 MPa, at 100 kg/h ratio (2020)
for 16 h
Apple pieces 35 ◦ C, 10 MPa, for 125 min Moisture content reduced to 18% •The shrinkage was more in ethanol SCCO2 and also Vetralla et al.
35 ◦ C, 10 MPa, for 50 min (40 Moisture content reduced to 11% had a wide pore size than raw and SCCO2 dried apple (2018)
min in Ethanol + SCCO2 for slices
10min
Coriander 40 ◦ C, 10 MPa, for 90 min 83% loss of moisture •SCD acts as both drying and microbial inactivation Zambon et al.
7

leaves Water activity reduced to 0.42 technique (2018)


Red pepper 40 ◦ C, 10 MPa for 16 h at a 100 90% loss of moisture •Increased TFC (20–40 to 880 mg/100 g) and •High retention of ascorbic acid and phenols than Vetralla et al.
kg/h flow ascorbic acid content (64–220 to 1163.20 mg/ fresh red pepper (2018)
100 g)
Beetroot 55 ◦ C, 60 MPa for 30 min •70%, 68%, and 42% retention of TPC, Marszałek et al.
betaxanthins, and betacyanins, respectively (2017)
a) Mango 40 C, 10 MPa for 300 min

a) 71% weight loss (aW = 0.124) •Carotenoid content decreased due to Braeuer et al.
b) Persimmon b) 84% weight loss (aW = 0.026) solubilization along with water (2017)
Basil leaves 40 ◦ C, 10 MPa, for 120 min 94% reduction in moisture content •93% retention of ascorbic acid •SCCO2 can be used as an alternate technique for Bušić et al.
•44% retention of TPC & TFC freeze-drying (2014)
•72% retention of chlorophyll
Apple slabs 45 ◦ C, 25 MPa Moisture content reduced to 1.5% •Better rehydration capacity, Wide pore distribution, Lee et al.
55 ◦ C, 25 MPa Moisture content reduced to 1.5% Caramelization of sugar (Burnt smell) (2011)
Carrot slices 50 ◦ C, 20 MPa for 150 min Increased moisture loss with increasing pressure and •Better rehydration ratio, shape, and less shrinkage Brown et al.
temperature (2008)

Journal of Food Engineering 343 (2023) 111375


TPC: Total phenolic content; TFC: Total flavonoid content; SCCO2: Supercritical CO2.
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Table 3
Influence of supercritical drying (SCD) on the rehydration capacity of food products.
Food Matrix Treatment conditions Drying rate Rehydration profile Key observations Reference

Chicken breast Rehydration: 30 C, 1:50 (w/v)



Morbiato
for 15 min et al. (2019)
a) SCCO2 40 C, 10 MPa for 420 min

Moisture removed ▪RC of 1.360 after 15 min ▪Micropore structure is significant in
was 70.9% aW: the combination method.
0.508
b) SCCO2 + High SCCO2: 40 ◦ C, 10 MPa for 420 min + Moisture removed ▪RC of 1.457 after 15 min ▪Greater retention of the original
power ultrasound HPU: 10 W for 300 min was 75.1% aW: <0.2 structure
c) Oven drying 75 ◦ C for 420 min Moisture removed ▪RC of 1.324 after 15 min ▪Poor rehydration, high degree of
was 74.4% aW: shrinkage,
0.453 ▪Case hardening is also the major
limitation
Carrot pieces Rehydration: 50 ◦ C or 80 ◦ C for Brown et al.
40 min (2008)
a) Modified SCCO2 Pre-treatment with 6% Ethanol at Moisture reduced to ▪Instant and high rehydration ▪Low density and porous structure
50 ◦ C for 40 min and with SCCO2 for 11% rate ▪Greater retention of original
10 min structure and shape

b) Pure SCCO2 50 ◦ C, 20 MPa for 150 min Moisture reduced to ▪Low porosity and less RC ▪Color loss was there, but it is
18% compared to ethanol-treated completely reversible after
sample rehydration
c) Air drying 40,50,60 ◦ C ▪Very poor retention capacity
▪A high degree of shrinkage with low
porosity
▪Destruction of interconnections
among the spaces of porous structure

SCD: Supercritical drying; SCF: Supercritical fluids; HPU: High-pressure ultra-sonication; SCCO2: Supercritical CO2; RC: Rehydration capacity.

4.3. Microbial and enzyme inactivation


Table 4
Modeling for rehydration kinetics for supercritical dehydrated products (Inyang
The efficiency of supercritical fluid to inactivate microbes and en­
et al., 2018; Marques et al., 2009).
zymes has been reported previously (Spilimbergo et al., 2018; Fer­
Model Equation Description rentino and Spilimbergo, 2011). A brief description is discussed for a
Peleg model
RR = RRe −
1
+
RR is the rehydration ratio, RRe is the better understanding of microbial and enzymatic inactivation (Fig. 4a
K2 saturation rehydration ratio, and k1 and and b). It occurs in a series of sequential steps (Hart et al., 2021; Lopes
t k2 are the Peleg constants.
K1 + K2 t et al., 2018; Ribeiro et al., 2020). Initially, the flushed cosolvent and SCF
First-order RR = RRe (RRe − 1) K3 is kinetic constant with dimension of solubilize in the external cell medium, which in turn reduces the
kinetics exp( − K3 t) time (min− 1) extracellular pH. In addition, the cosolvent may cause the desiccation of
model
cells. The diffusion of CO2 into the cell decreases the intracellular pH.
Weibull model RR = RRe (1 − RRe ) α is the scale parameter, and β is the shape
factor
The primary enzyme system becomes compromised at a modified pH;
[ ( t )β ]
exp − the enzymes unfold & denature and lose their capability to participate in
α
metabolic activities (Hart et al., 2021) (Fig. 4b). The metabolism
disturbance is due to CO2 and HCO3 – ion formation. The balance of
standard error of estimation (SEE), and mean relative percent deviation intracellular electrolytes is disturbed due to the binding of CO2− 3 with
(MRD) to predict the rehydration kinetics of dried food samples (Mar­ Ca2+ & Mg2+ binding proteins and inorganic electrolytes. Therefore, the
ques et al., 2009). SEE and MRD are expressed in Eqs (6) and (7) as phospholipids and hydrophobic compounds of cell walls and mem­
follows: branes become solubilized. Since the metabolism of microbes is
(∑( )2 )0.5 completely disturbed, they lose their viability. The extent of microbial
RRn,exp − RRn,cal inactivation after SCD of various commodities is summarized in Table 5.
SEE = (6)
df The microbial reduction is more when high moisture is present in food
(Bourdoux et al., 2018). As a result, more work is needed to back up the
⃒ ⃒
100 ∑ ⃒RRn,exp − RRn,cal ⃒ claim that a more substantial biocide effect was observed at higher water
MRD = (7) activity. One may maintain a lower rate of drying at the beginning of the
N RRn,exp
drying process to ensure that the water activity is high, which helps
Here RRn,exp is the experimental rehydration ratio, RRn,cal is the increase the inactivation of microbes (Liu et al., 2018). Zambon et al.
theoretical rehydration ratio, N is the number of data obtained, and df is (2021) conducted a study on apple slices at 40 ◦ C/10 MPa/10 min and
the number of degrees of freedom of the regression model (number of then depressurized for 20 min reduced E. coli O157:H7, Salmonella
data points – number of constants in the model). Higher R2 values and typhimurium SL 1344, and Listeria monocytogenes from 5.56, 5.63, and
lower SEE and MRD values of the model indicate it fits the experimental 7.41 log CFU/g to 1 log CFU/g, respectively. The increase in the number
data. MRD values < 10% are indicative of the best fit for many practical of depressurization steps led to higher microbial and enzyme inactiva­
purposes (Marques et al., 2009). The prediction models have not yet tion due to the sudden difference in pressure (Wimmer and Zarevúcka,
been applied to the rehydration kinetics of supercritical dried samples. 2010). SCD of coriander at 35 ◦ C/8 MPa/150 min reduced the aerobic
Hence, no kinetic modeling has been developed for supercritical dried mesophiles, yeasts & molds, and Enterobacteriaceae by 2.80, 5.03, and
samples, indicating a need for more research in this area for the vali­ 4.61 log cycles, respectively, and was far better than freeze drying,
dation of the data. which showed 1.23, 0.87, and 0.97 log cycles reduction, respectively.
The mesophilic spores did not reduce significantly after SCD (0.11 log
cycle reduction). The log reduction of Salmonella, E. coli O157:H7, and

8
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Fig. 4. A schematic diagram showing the possible mechanism of (a) microbial and (b) enzyme inactivation during the supercritical drying process.

L. monocytogenes was 5.29, 5.69 and 5.55, respectively (Bourdoux et al., for a year. SCCO2 dried apples packed in nitrogen-flushed alumi­
2018). A similar trend was observed in coriander by Michelino et al. nium-polyethylene bags maintained good sensory acceptability and
(2018). Additional research is needed to assess its potential to inactivate nutritional quality. Tomic et al. (2020) conducted a comparative study
various microorganisms and their species. on supercritical dried (40 ◦ C/10 MPa/14 h), freeze-dried
Enzymes are biological catalysts, which are mostly proteins. The (− 25 ◦ C/0.00002 MPa/48 h), and fried (160 ◦ C for 3.5 min) beetroot
enzyme activity is reduced by increasing CO2 pressure (Hart et al., 2021) for sensory analysis and overall acceptability. They found that SCD
and is shown to be completely lost at pressures above 30 MPa (Rezaei without precooking had exceptional sensory and overall acceptability.
et al., 2007). There are possibilities for the formation of carbamates by However, an economic justification is necessary for large-scale com­
the reaction of CO2 with the amino group of the enzyme or protein. A mercial use. The SCD samples can achieve better sensory, quality, and
recent study performed by Zambon et al. (2022b) on strawberries consumer approval than freeze-drying and fried samples. A similar study
indicated complete inactivation of PPO and POD under treatment con­ on carrots was conducted by Brown et al. (2008), which corroborated
ditions of 40 ◦ C/13.3 MPa/420 min at a 19 kg/h flow rate. No other the same trend. At 40 ◦ C/14 MPa/160 kg/h/16 h, red bell pepper had
information is available on the effect of SCD on the inactivation of en­ lower levels of hardness and cohesiveness than freeze-dried samples and
zymes. No general assumptions can be drawn on this aspect due to the air-dried samples. The SC dried red bell pepper was accepted by >60%
scarcity of available information. Hence, more research efforts on mi­ of the sensory-tested consumers. During storage, the samples lost their
crobial and enzyme inactivation by SCD are obligatory. sweet pepper flavor and developed a hay-like odor (Zambon et al.,
2020).
4.4. Organoleptic properties
4.5. Color profile
The parameters considered for the evaluation of organoleptic prop­
erties vary significantly for different products. Tomic et al. (2019) Color is one of the most important factors that draw the attention of
evaluated the change in sensory properties and secondary metabolites in customers at first glance. Conventional drying has adverse effects on
SCCO2 dried ‘Elstar’ apple cuts and snacks stored at room temperature retaining the color of the dried samples. Therefore, it is essential to know

9
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Table 5 Table 6
Influence of SCD on the microbial decontamination of food products. Influence of supercritical drying on the color profile of food products.
Food Matrix Treatment Target Max log References Food Treatment Max changes in Key observations References
conditions microorganism reduction Matrix conditions color profile
obtained
Red bell 40 ◦ C, 14 MPa ▪Orange color ▪The loss of Zambon
Strawberry 40 ◦ C, 13.3 [1]E. coli O157:H7 [1] >6 Zambon pepper with 160 kg/h (red color loss) carotene due to et al.
MPa for 420 [2] Salmonella Sps [2] >6 et al. its lipophilic (2020)
min at 19 kg/h [3] Listeria [3] >5 (2022b) nature and
monocytogenes dissolution in
Strawberry 40 ◦ C, 10 MPa [1] E. coli O157: [1] 5 Zambon SCCO2
slices for 360 min H7 et al. Apple 35 ◦ C, 10 MPa ▪No overall ▪An increase in Vetralla
[2] Salmonella Sps [2] 5 (2022a) for 125 min change in the L*, a*, and b* et al.
[3] Listeria [3] 5 35 ◦ C, 10 MPa color values were (2018)
monocytogenes for 50 min observed only in
Apple slices 40 ◦ C, 10 MPa [1] Listeria [1] 6.41 Zambon (40 min in SCCO2 dried
for 10 min monocytogenes et al. ethanol + apples.
[2] Salmonella. [2] 4.42 (2021) 10min in
Typhimurium SCCO2)
[3] E. coli O157: [3] 4.56 Coriander US + SCD ▪The original ▪L * and a* value Michelino
H7 40 W, 40 ◦ C, color was increased, b* et al.
Chicken [a] SCD [1.a] Mesophilic [1.a] >6 Morbiato 10 MPa for 90 maintained value decreased, (2018)
breast 40 ◦ C, 10 MPa bacteria et al. min and a slight
for 420 min [2.a] Salmonella [2.a] >7 (2019) change in overall
[b] US + SCD enterica color.
HPU: 10 W for [3.a] Yeast and [3.a] 3.88 Basil 40 ◦ C, 10 MPa ▪Better ▪Lower L * and Bušić et al.
300 min molds leaves for 180 min at retention of b* and higher a* (2014)
SCCO2: 40 ◦ C, [1.b] Mesophilic [1.b] >6 1000 L/h green color value than
10 MPa for bacteria compared to frozen and fresh
420 min [2.b] Salmonella [2.b] >7 air-dried basil leaves
enterica sample (wilted
[3.b] Yeast and [3.b] 3.88 and browned)
molds ▪Darker in
Coriander 35 C, 8 MPa

[1] E. coli O157: [1] 5.29 Bourdoux color compared
for 150 min H7 et al. to fresh Basil
[2] Salmonella sps. [2] 5.69 (2018) leaves
[3] Aerobic [3] 2.80 Carrot 40–60 ◦ C, 20 ▪Pale in color ▪The color of the Brown
mesophiles MPa for 30 than raw carrot rehydrated et al.
[4] Mesophilic [4] 0.11 min after drying sample was like (2008)
spores 6% Ethanol, that of the raw
[5] [5] 4.61 40–60 ◦ C, 20 sample,
Enterobacteriaceae MPa for 30 indicating that
[6] Yeasts and [6] >5.0 min no loss of
molds β-carotene
Coriander [a] SCD [1.a] Mesophilic [1.a] 3.63 Michelino
40 ◦ C, 10 MPa bacteria et al. SCF: supercritical fluids; SCD: supercritical drying; US: ultrasonication;.
for 90 min [2.a] Mesophilic [2.a] 0.2 (2018)
[b] US + SCD spores deposition of some of the carotenes within a specific area of the sample
40 W, 40 ◦ C, [3.a] Yeast and [3.a] 3.57
(Brown et al., 2008). This contrasts with the study by Zambon et al.
10 MPa for 90 molds
min [1.b] Mesophilic [1.b] 4.8 (2020), who showed that the color of red bell peppers treated at
bacteria 40 ◦ C/14 MPa/160 kg/h was lost due to the dissolution of carotene in
[2.b] Mesophilic [2.b] 1.7 SCCO2 when compared to freeze drying. SCD of basil leaves at 40 ◦ C/10
spores
MPa/1000 L/h/3 h has shown a loss of Mg+2, which was responsible for
[3.b] Yeast and [3.b] 3.57
molds
the drastic loss of color, which changed from greenish blue to greenish
Coriander 40 ◦ C, 10 MPa [1] Mesophilic [1] 4 Zambon yellow (Bušić et al., 2014). It is observed that the color of a few products
leaves for 150 min bacteria et al. is affected to a greater extent by SCD due to its non-polar nature. In­
[2] Mesophilic [2] <1 (2018) vestigations on the impact of SCD on color and its mechanism need
spores
attention from researchers.
[3] Yeast and [3] >5
molds
4.6. Shelf life
SCF: supercritical fluids; SCD: supercritical drying; US: ultrasonication; Ethanol.

Shelf life studies are carried out on food products to project the end-
the effect of SCD on the color of products and subsequently arrive at
of-life span and to indicate the ‘best before’ dates for consumer safety
optimum processing and storage conditions. A summary of the studies
and awareness. SCD products have a comparatively higher shelf life than
exploring the changes in color profile in the SCD samples is presented in
those dried using various other drying techniques. Literature on these
Table 6. Sometimes, SCD may lead to color loss due to the high solubility
aspects is scant, and these have been summarized in Table 7. Djekic et al.
of pigments and volatiles in SCF. When treated with SCCO2 (50 ◦ C/20
(2018) studied the effects of SCD, air drying, and freeze drying on apples
MPa/150 min) and ethanol-SCCO2 (6% ethanol (40 min) + 50 ◦ C/20
for six months under ambient conditions and measured different quality
MPa/10 min), the color of SCD carrot slices became pale compared to
parameters (texture, color, and sensory properties). Apples dried by SCD
air-dried and control. This is because of the higher solubility of carot­
showed the highest consumer acceptability at the beginning of the
enoids in SCCO2, which might get lost on SCD. However, the color was
study. After six months, the SCD and freeze-dried apples packed in
similar to the original samples on rehydration which suggests a mini­
nitrogen-flushed polyethylene-covered aluminium showed no signifi­
mum loss of carotenoids during the SCD process. The air-dried sample
cant color, texture, or sensory changes. Both processes yielded almost
was darker due to enzymatic browning, damaged cell structure, and the
similar results. When apples were treated with SCD at 50 ◦ C/12.5

10
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Table 7
Changes in quality attributes during storage of supercritical dried food products.
Food Treatment Packaging material Organoleptic properties Storage Max changes in the Key observations Reference
Matrix conditions analyzed conditions key property after
(Shelf-life) shelf-life

Apple cuts/ 50 ◦ C, 12.5 Aluminium- 12 months High sensory and ▪High preservation of Tomic et al.
snacks MPa for 16 h polyethylene bags nutritional quality flavanols (2019)
flushed with N2
Apple 50 ◦ C, 12.5 Aluminium- Good product springiness 6 months The color was ▪Flavor and aroma were Djekic et al.
(Elstar) MPa for 16 h polyethylene bags with better water activity changed to greyish changed with hay odor (2018)
flushed with N2 and nutrient retention brown ▪Increased chewiness with
loss of crispiness and
flexibility
▪Increased oxidation
Coriander a) SCD An aluminium bag 10 days at The original color ▪ Microbial growth Michelino
40 ◦ C, 10 MPa filled with 100% N2 30 ◦ C was maintained. was minimum; water et al. (2018)
for 90 min loss was consistent
b) US + SCD
40 W, 40 ◦ C,
10 MPa for 90
min

SCF: Supercritical fluids; SCD: Supercritical drying; US: Ultrasonication.

MPa/16 h and freeze-dried at − 25 ◦ C/0.00002MPa/24 h, the shelf life of gathers both energy requirements and fluid costs. The energy utility is
SC-dried apples was extended to one year under refrigerated storage comparatively less expensive because the heat is generated through
conditions (4 ◦ C). The average flavonoid content was similar for both steam or hot water using fuel or gas for installation in small and
samples during the storage period. The total quality index was best for medium-scale industries. Electrical heating can also be used but is
SCD-dried apples than freeze-dried samples, with scores of 0.39 and limited to small-scale sectors (Lázár and Fábián, 2016). The consump­
0.44, respectively (Tomic et al., 2020). tion of CO2 is a significant factor in estimating the variable cost. In the
The loss of triterpenes, dihydrochalcones, hydroxycinnamic acid, SCD process, depressurization may cause loss of CO2 and organic
flavanols, and proanthocyanins is more significant in SCD apple slices cosolvent. Hence, recompression may compensate for the CO2 loss on a
during the storage period than in the freeze-dried sample. This study commercial scale. Further recycling of CO2 is another cost-saving option
opined that freeze drying is superior in preserving the secondary me­ (Lázár and Fábián, 2016). In most studies, the processing time of the
tabolites than SCD; however, lower concentrations of bioactives were complete SCD process varies from 30 min to 7 h based on the food type.
reported after SCD treatment at 50 ◦ C. By keeping the SCD process The sanitation process of SCD equipment is also essential, but it is
temperature below 45 ◦ C, higher retention of bioactive components can consistently underestimated in the cost estimation. Therefore, it is
be maintained with nutritional quality similar to the freeze-dried crucial to include the cleaning issue at the beginning of the process.
product (Tomic et al., 2019). Nevertheless, additional studies on From the product perspective, high-value products such as spices and
different process conditions should be performed to confirm this condiments, meat products, and aerogels are being treated with SCD,
hypothesis. which has a high market value. Hence, retention of its originality is
highly expected, and previous literature does indicate that SCD fulfils
5. Economic and environmental aspects of SCD this requirement. The limitations of supercritical technology are the
higher installation cost and maintenance due to the involvement of both
There are speculations on the cost of SCD because of its higher in­ temperature and pressure, the use of cosolvent reservoirs, and the
vestment cost than other conventional drying techniques which are establishment and maintenance of predetermined processing conditions.
currently used commercially. There are only 2 to 3 articles on the eco­ Hence, the cost of supercritical-based processes is relatively higher than
nomics of SCD; based on the information available, this section is con­ other conventional methods.
structed. SCD installation depends on the equipment, instrumentation, The characteristics of the SCD process that relate to environmental
and automation. A typical SCD process includes drying reactor vessels, impact are sustainability, reduced water requirements for cleaning and
liquid CO2 reservoirs, and a pumping facility at a set pressure. A heat processing, decreased effluent burden, and reduced energy use. As it is
exchanger is required to bring the desired temperature of SCF and carried out in a totally closed system, no additional facility is needed for
maintain the temperature of the system. In addition, piping, valves, discharged wastewater or exhaust gases. It does not involve the release
instruments, and other safe automation and operation services will add of processed, harmful substances such as heavy metals, halogenated
to the initial investment cost. The flow rate of solvent, Q, is directly carbons, acrylamide, or polycyclic aromatic hydrocarbons, unlike other
proportional to the volume of the extractors (VT); this indicates that the drying processes. Without harming the environment, the SCF used in the
investment cost for SCD increases by its capacity. This relationship in­ SCD process is recycled and put to other uses. It is expected that SCD
dicates that an increase in the volume of the drying chamber or flow rate could reduce pollution and contribute substantially to improving envi­
of the SCF decreases the cost of the SCD process exponentially. It can be ronmental sustainability in the forthcoming coming decades.
applied from a pilot scale of 0.5 L capacity to a commercial scale of 500 L
capacity. Nevertheless, it underestimates the prices of small or pilot 6. Challenges and future scope
scales and overestimates the prices of commercial units, which are likely
to be lower by this rule. SCD has been carried out on only a few foods, such as apples, carrots,
SCD can be performed in both batch and continuous processes beetroot, coriander, red bell pepper, and chicken breast. Even though a
(workforce costs reduced). The continuous process of SCD can be few researchers have worked on the microbial inactivation of SCD, there
decreased to one-third of the batch process in terms of solvent con­ is a scarcity of complete information on the inactivation induced by the
sumption. In contrast, there is a prominent increase in the price when SCD process. Most of the studies have focused on a few bacterial species,
the cosolvent-mediated SCD process is employed with a different yeasts, and molds species, which cannot be extrapolated to destroy all
implementation of the ethanol reservoir. The variable cost of the SCD kinds of microorganisms. Hence, research should focus on a broader

11
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

range of foods and microorganisms for a specific strain because of Data availability
changes in the specificity of the microorganisms to different SCD
treatment conditions, as is the case with other drying techniques. One of No data was used for the research described in the article.
the major challenges during the SCD process is the loss of other polar
soluble compounds along with water. There is no validated study in this Acknowledgments
regard and whether modifying the SCD conditions can control the loss of
polar compounds and specifically remove water. This research did not receive any specific grant from funding
It is essential to study SCD in combination with different cosolvents agencies in the public, commercial, or not-for-profit sectors.
to maintain the product’s structure and gather information to under­
stand the mechanisms of different foods and phase behavior before References
coming to general assumptions. It is imperative to understand the drying
kinetics of SCD, the effect of temperature and pressure combinations, Asafu-Adjaye, O., Via, B., Sastri, B., Banerjee, S., 2021. Mechanism of dewatering porous
structures with supercritical carbon dioxide. Case Studies in Chemical and
and the processing time on different food components. These are Environmental Engineering 4, 100128. https://doi.org/10.1016/j.cscee.2021.100
generally performed as batch types, although continuous processes are 128.
also possible. A few moderate-cost drying techniques such as SCD fall Batiha, G.E.S., Hussein, D.E., Algammal, A.M., George, T.T., Jeandet, P., Al-Snafi, A.E.,
Cruz-Martins, N., 2021. Application of natural antimicrobials in food preservation:
between these two extremes. More information is required for further recent views. Food Control 126, 108066. https://doi.org/10.1016/j.
evaluation on the delivery of high quality with good overall accept­ foodcont.2021.108066.
ability, microbial safety, enzyme inhibition, physicochemical changes, Benali, M., Canada, N.R., Boumghar, Y., 2014. Energy aspects in drying. Handbook of
Industrial Drying, July 1107–1130. https://doi.org/10.1201/b17208-68.
bioactive compounds, and other quality changes during shelf life studies Bourdoux, S., Rajkovic, A., De Sutter, S., Vermeulen, A., Spilimbergo, S., Zambon, A.,
of the food to ensure that the food is safe for consumption. Hofland, G., Uyttendaele, M., Devlieghere, F., 2018. Inactivation of Salmonella,
Listeria monocytogenes, and Escherichia coli O157:H7 inoculated on coriander by
freeze-drying and supercritical CO2 drying. Innovat. Food Sci. Emerg. Technol. 47,
7. Conclusion
180–186. https://doi.org/10.1016/j.ifset.2018.02.007.
Braeuer, A.S., Schuster, J.J., Gebrekidan, M.T., Bahr, L., Michelino, F., Zambon, A.,
Supercritical drying (SCD) has emerged as a sustainable and prom­ Spilimbergo, S., 2017. In situ Raman analysis of CO2-assisted drying of fruit slices.
ising technique for food commodities such as spices, seafood, a few fruits Foods 6 (5), 1–11. https://doi.org/10.3390/foods6050037.
Brown, Z.K., Fryer, P.J., Norton, I.T., Bakalis, S., Bridson, R.H., 2008. Drying of foods
and vegetables, and other semisolids such as agar gels and barley using supercritical carbon dioxide — investigations with carrot. Innovat. Food Sci.
β-glucan gels. The SCD process removes water from foods using SCF Emerg. Technol. 9 (3), 280–289. https://doi.org/10.1016/J.IFSET.2007.07.003.
without crossing the liquid-vapor boundary layer while maintaining Brunner, G., 2005. Supercritical fluids: technology and application to food processing.
J. Food Eng. 67 (1–2), 21–33. https://doi.org/10.1016/J.JFOODENG.2004.05.060.
combined critical temperature and pressure conditions. The combina­ Bušić, A., Vojvodić, A., Komes, D., Akkermans, C., Belščak-Cvitanović, A., Stolk, M.,
tion of CO2 and ethanol in a supercritical state is being used to improve Hofland, G., 2014. Comparative evaluation of CO2 drying as an alternative drying
drying efficiency as these are non-toxic, accessible, affordable, and eco- technique of basil (Ocimum basilicum L.) - the effect on bioactive and sensory
properties. Food Res. Int. 64, 34–42. https://doi.org/10.1016/j.
friendly. Retention of physical, chemical, microbial, and sensory prop­ foodres.2014.06.013.
erties of food commodities is very high in SCD-treated samples. It pre­ Chakraborty, S., Shaik, L., Gokhale, J.S., 2020. Subcritical Water: an Innovative
serves thermally sensitive components of the food to the maximum Processing Technology. Innovative Food Processing Technologies: A Comprehensive
Review. Elsevier. https://doi.org/10.1016/b978-0-08-100596-5.22966-1.
while retaining the microstructure of the products. SCF can be recycled Comin, L.M., Temelli, F., Saldaña, M.D.A., 2012. Barley beta-glucan aerogels via
and reused for other drying cycles without polluting the environment. In supercritical CO2 drying. Food Res. Int. 48 (2), 442–448. https://doi.org/10.1016/j.
a few food products, SCD does not guarantee color retention, which is foodres.2012.05.002.
Corzzini, S.C.S., Barros, H.D.F.Q., Grimaldi, R., Cabral, F.A., 2017. Extraction of edible
essential in most foods. There is a chance to widen the range of products
avocado oil using supercritical CO2 and a CO2/ethanol mixture as solvents. J. Food
and microbes. More focus on the bioactive, microbial, and enzyme Eng. 194, 40–45. https://doi.org/10.1016/j.jfoodeng.2016.09.004.
inactivation and shelf life of supercritical dried foods is needed. There is Derossi, A., De Pilli, T., Severini, C., McCarthy, M.J., 2008. Mass transfer during osmotic
scope for research in determining the drying and rehydration kinetics, dehydration of apples. J. Food Eng. 86 (4), 519–528. https://doi.org/10.1016/j.
jfoodeng.2007.11.007.
computational fluid dynamics, and modeling of the SCD process, which Djekic, I., Tomic, N., Bourdoux, S., Spilimbergo, S., Smigic, N., Udovicki, B., Hofland, G.,
is essential to design a process at the commercial level. Devlieghere, F., Rajkovic, A., 2018. Comparison of three types of drying
(supercritical CO2, air, and freeze) on the quality of dried apple – quality index
approach. Lebensm. Wiss. Technol. 94 (November 2017), 64–72. https://doi.org/
Author contributions 10.1016/j.lwt.2018.04.029.
Ferrentino, G., Spilimbergo, S., 2011. High-pressure carbon dioxide pasteurization of
Sourcing literature, drafting the original article, preparing the fig­ solid foods: current knowledge and future outlooks. Trends Food Sci. Technol. 22
(8), 427–441. https://doi.org/10.1016/j.tifs.2011.04.009.
ures, preparing the tables: Kosana Pravallika; Design of review outline, García-Pascual, P., Sanjuán, N., Melis, R., Mulet, A., 2006. Morchella esculenta (morel)
supervising the research, reviewing and revising the manuscript, editing rehydration process modelling. J. Food Eng. 72 (4), 346–353.
the manuscript: Snehasis Chakraborty; The conception of the idea, Girotra, P., Singh, S.K., Nagpal, K., 2013. Supercritical fluid technology: A promising
approach in pharmaceutical research. Pharmaceutical Development and Technology
design of review outline, reviewing and revising the manuscript: Rekha 18 (1), 22–38. https://doi.org/10.3109/10837450.2012.726998.
S. Singhal. Harrison, R., 2020. Fluids and the field. Elements 16 (6), , 367. https://doi.org/10.2138/
GSELEMENTS.16.6.430.
Hart, A., Anumudu, C., Onyeaka, H., Miri, T., 2021. Application of supercritical fluid
Ethical approval
carbon dioxide in improving food shelf life and safety by inactivating spores: a
review. Journal of Food Science and Technology. https://doi.
This work does not contain any studies with human participants or org/10.1007/s13197-021-05022-7.
animals performed by any of the authors. Inyang, U.E., Oboh, I.O., Etuk, B.R., 2018. Kinetic Models for drying techniques—food
materials. Adv. Chem. Eng. Sci. 8 (2), 27–48. https://doi.org/10.4236/
aces.2018.82003.
Declaration of competing interest Khaw, K.Y., Parat, M.O., Shaw, P.N., Falconer, J.R., 2017. Solvent supercritical fluid
technologies to extract bioactive compounds from natural sources: a review.
Molecules 22 (7). https://doi.org/10.3390/molecules22071186.
The authors declare that they have no known competing financial Krokida, M.K., Marinos-Kouris, D., 2003. Rehydration kinetics of dehydrated products.
interests or personal relationships that could have appeared to influence J. Food Eng. 57 (1), 1–7.
the work reported in this paper. Lázár, I., Fábián, I., 2016. A continuous extraction and pumpless supercritical CO2 drying
system for laboratory-scale aerogel production. Gels 2 (4). https://doi.org/10.3390/
gels2040026.

12
K. Pravallika et al. Journal of Food Engineering 343 (2023) 111375

Lee, B.S., Choi, Y.H., Lee, W.Y., 2011. Drying characteristics of apple slabs after pre- Spilimbergo, S., Matthews, M.A., Cinquemani, C., 2018. Supercritical Fluid
treatment with supercritical CO2. Journal of Food Science and Nutrition 16 (3), Pasteurization and Food Safety. In: Proctor, A. (Ed.), Alternatives to Conventional
261–266. https://doi.org/10.3746/jfn.2011.16.3.261. Food Processing. Royal Society of Chemistry, UK, pp. 153–195.
Lee, L.Y., Lin, Y., Wang, C.-H., 2016. Study of the supercritical drying of wet okara. July Suravanichnirachorn, W., Haruthaithanasan, V., Suwonsichon, S., Sukatta, U.,
1–7. In: http://www.itekcmsonline.com/rps2prod/set2016/usb-proceedings/html/ Maneeboon, T., Chantrapornchai, W., 2018. Effect of carrier type and concentration
P207.xml. on the properties, anthocyanins and antioxidant activity of freeze-dried mao
Lewicki, P.P., 2004. Water as the determinant of food engineering properties. A review. [Antidesmabunius L. Spreng] powders. Agriculture and Natural Resources 52 (4),
J. Food Eng. 61 (4), 483–495. https://doi.org/10.1016/S0260-8774(03)00219-X. 354–360. https://doi.org/10.1016/j.anres.2018.09.011.
Liu, S., Ozturk, S., Xu, J., Kong, F., Gray, P., Zhu, M.J., Tang, J., 2018. Microbial Thorat, I.D., Mohapatra, D., Sutar, R.F., Kapdi, S.S., Jagtap, DipaliD., 2012.
validation of radio frequency pasteurization of wheat flour by inoculated pack Mathematical modeling and experimental study on thin layer vacuum drying of
studies. J. Food Eng. 217, 68–74. https://doi.org/10.1016/j.jfoodeng.2017.08.013. ginger (Zingiber officinale R.) slices. Food Bioprocess Technol. 5, 1379–1383.
Lopes, R.P., Mota, M.J., Gomes, A.M., Delgadillo, I., Saraiva, J.A., 2018. Application of Tomic, N., Djekic, I., Hofland, G., Smigic, N., Udovicki, B., Rajkovic, A., 2020.
high pressure with homogenization, temperature, carbon dioxide, and cold plasma Comparison of supercritical CO2-drying, freeze-drying and frying on sensory
for the inactivation of bacterial spores: a review. Compr. Rev. Food Sci. Food Saf. 17 properties of beetroot. Foods 9 (9). https://doi.org/10.3390/foods9091201.
(3), 532–555. https://doi.org/10.1111/1541-4337.12311. Tomic, N., Djekic, I., Zambon, A., Spilimbergo, S., Bourdoux, S., Holtze, E., Hofland, G.,
Lopez-Quiroga, E., Prosapio, V., Fryer, P.J., Norton, I.T., Bakalis, S., 2019. A model-based Sut, S., Dall’Acqua, S., Smigic, N., Udovicki, B., Rajkovic, A., 2019. Challenging
study of rehydration kinetics in freeze-dried tomatoes. Energy Proc. 161, 75–82. chemical and quality changes of supercritical CO2 dried apple during long-term
Marszałek, K., Krzyżanowska, J., Woźniak, Ł., Skąpska, S., 2017. Kinetic modelling of storage. Lebensm. Wiss. Technol. 110, 132–141. https://doi.org/10.1016/j.
polyphenol oxidase, peroxidase, pectin esterase, polygalacturonase, degradation of lwt.2019.04.083.
the main pigments and polyphenols in beetroot juice during high pressure carbon Uwineza, P.A., Waśkiewicz, A., 2020. Recent advances in supercritical fluid extraction of
dioxide treatment. Lebensm. Wiss. Technol. 85, 412–417. https://doi.org/10.1016/ natural bioactive compounds from natural plant materials. Molecules 25 (17).
j.lwt.2016.11.018. https://doi.org/10.3390/molecules25173847.
Marques, L.G., Prado, M.M., Freire, J.T., 2009. Rehydration characteristics of freeze- Vatai, T., Škerget, M., Knez, Ž., 2009. Extraction of phenolic compounds from elder berry
dried tropical fruits. LWT–Food Sci. Technol. 42 (7), 1232–1237. and different grape marc varieties using organic solvents and/or supercritical carbon
Michelino, F., Zambon, A., Vizzotto, M.T., Cozzi, S., Spilimbergo, S., 2018. High power dioxide. J. Food Eng. 90 (2), 246–254. https://doi.org/10.1016/j.
ultrasound combined with supercritical carbon dioxide for the drying and microbial jfoodeng.2008.06.028.
inactivation of coriander. J. CO2 Util. 24, 516–521. https://doi.org/10.1016/J. Vetralla, M., Ferrentino, G., Zambon, A., Spilimbergo, S., 2018. A study about the effects
JCOU.2018.02.010. of supercritical carbon dioxide drying on apple pieces. ETP International Journal of
Morbiato, G., Zambon, A., Toffoletto, M., Poloniato, G., Dall’Acqua, S., de Bernard, M., Food Engineering 4 (3), 186–190. https://doi.org/10.18178/ijfe.4.3.186-190.
Spilimbergo, S., 2019. Supercritical carbon dioxide combined with high power Wang, W., Rao, L., Wu, X., Wang, Y., Zhao, L., Liao, X., 2021. Supercritical Carbon
ultrasound as innovate drying process for chicken breast. J. Supercrit. Fluids 147, Dioxide Applications in Food Processing. Food Engineering Reviews 13 (3),
24–32. https://doi.org/10.1016/J.SUPFLU.2019.02.004. 570–591. https://doi.org/10.1007/s12393-020-09270-9.
Moses, J.A., Norton, T., Alagusundaram, K., Tiwari, B.K., 2014. Novel drying techniques Wason, S., Verma, T., Subbiah, J., 2021. Validation of process technologies for enhancing
for the food industry. Food Eng. Rev. 6 (3), 43–55. https://doi.org/10.1007/s12393- the safety of low-moisture foods: a review. Compr. Rev. Food Sci. Food Saf. 20 (5),
014-9078-7. 4950–4992. https://doi.org/10.1111/1541-4337.12800.
Nautiyal, O.H., 2016. Food processing by supercritical carbon dioxide-Review. EC Wimmer, Z., Zarevúcka, M., 2010. A review on the effects of supercritical carbon dioxide
Chemistry 2, 111–135. on enzyme activity. Int. J. Mol. Sci. 11 (1), 233–253. https://doi.org/10.3390/
Nowacka, M., Dadan, M., Janowicz, M., Wiktor, A., Witrowa-Rajchert, D., Mandal, R., ijms11010233.
Pratap-Singh, A., Janiszewska-Turak, E., 2021. Effect of nonthermal treatments on Wrona, O., Rafińska, K., Możeński, C., Buszewski, B., 2017. Supercritical fluid extraction
selected natural food pigments and colour changes in plant material. Compr. Rev. of bioactive compounds from plant materials. Journal of AOAC International, 100
Food Sci. Food Saf. 20 (5), 5097–5144. https://doi.org/10.1111/1541-4337.12824. (6), 1624-1635. https://doi.org/10.5740/jaoacint.17-0232.
Phisut, N., 2012. Spray drying technique of fruit juice powder: some factors influencing Zambon, A., Facco, P., Morbiato, G., Toffoletto, M., Poloniato, G., Sut, S., Andrigo, P.,
the properties of product. International Food Research Journal 19 (1), 7–18. Dall’Acqua, S., de Bernard, M., Spilimbergo, S., 2022b. Promoting the preservation
Plazzotta, S., Calligaris, S., Manzocco, L., 2019. Structure of oleogels from κ-carrageenan of strawberry by supercritical CO2 drying. Food Chemistry, 397(October 2021),
templates as affected by supercritical-CO2 -drying, freeze-drying and lettuce-filler 133789. https://doi.org/10.1016/j.foodchem.2022.133789.
addition. Food Hydrocolloids 96 (March), 1–10. https://doi.org/10.1016/j. Zambon, A., Bourdoux, S., Pantano, M.F., Pugno, N.M., Boldrin, F., Hofland, G.,
foodhyd.2019.05.008. Rajkovic, A., Devlieghere, F., Spilimbergo, S., 2021. Supercritical CO2 for the drying
Rezaei, K., Temelli, F., Jenab, E., 2007. Effects of pressure and temperature on enzymatic and microbial inactivation of apple’s slices. Dry. Technol. 39 (2), 259–267. https://
reactions in supercritical fluids. Biotechnol. Adv. 25 (3), 272–280. https://doi.org/ doi.org/10.1080/07373937.2019.1676774.
10.1016/j.biotechadv.2006.12.002. Zambon, A., Michelino, F., Bourdoux, S., Devlieghere, F., Sut, S., Dall’Acqua, S.,
Ribeiro, N., Soares, G.C., Santos-Rosales, V., Concheiro, A., Alvarez-Lorenzo, C., García- Rajkovic, A., Spilimbergo, S., 2018. Microbial inactivation efficiency of supercritical
González, C.A., Oliveira, A.L., 2020. A new era for sterilization based on supercritical CO2 drying process. Dry. Technol. 36 (16), 2016–2021. https://doi.org/10.1080/
CO2 technology. J. Biomed. Mater. Res., Part B 108 (2), 399–428. https://doi.org/ 07373937.2018.1433683.
10.1002/jbm.b.34398. Zambon, A., Tomic, N., Djekic, I., Hofland, G., Rajkovic, A., Spilimbergo, S., 2020.
Şahin, İ., Özbakır, Y., İnönü, Z., Ulker, Z., Erkey, C., 2018. Kinetics of supercritical drying Supercritical CO2 drying of red bell Pepper. Food Bioprocess Technol. 13 (5),
of gels. Gels 4 (1). https://doi.org/10.3390/gels4010003. 753–763. https://doi.org/10.1007/s11947-020-02432-x.
SalimiHizaji, A., Maghsoudlou, Y., Jafari, S.M., 2010. Application of Peleg model to Zambon, A., Zulli, R., Boldrin, F., Spilimbergo, S., 2022a. Microbial inactivation and
study effect of water temperature and storage time on rehydration kinetics of air drying of strawberry slices by supercritical CO2. J. Supercrit. Fluids 180, 105430.
dried potato cubes. Lat. Am. Appl. Res. 40 (2), 131–136. https://doi.org/10.1016/J.SUPFLU.2021.105430.
Santhalakshmy, S., Bosco, S.J.D., Francis, S., Sabeena, M., 2015. Effect of inlet Zheng, S., Hu, X., Ibrahim, A.R., Tang, D., Tan, Y., Li, J., 2010. Supercritical fluid drying:
temperature on physicochemical properties of spray-dried jamun fruit juice powder. classification and applications. Recent Pat. Chem. Eng. 3 (3), 230–244. https://doi.
Powder Technol. 274, 37–43. https://doi.org/10.1016/j.powtec.2015.01.016. org/10.2174/1874478811003030230.

13

You might also like