Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Trends in Food Science & Technology 63 (2017) 1e17

Contents lists available at ScienceDirect

Trends in Food Science & Technology


journal homepage: http://www.journals.elsevier.com/trends-in-food-science-
and-technology

Review

Spray drying of probiotics and other food-grade bacteria: A review


nae
Song Huang a, b, Marie-Laure Vignolles c, Xiao Dong Chen a, Yves Le Loir b, Gwe €l Jan b,
b a, b, *
Pierre Schuck , Romain Jeantet
a
Suzhou Key Lab of Green Chemical Engineering, School of Chemical and Environmental Engineering, College of Chemistry, Chemical Engineering and
Material Science, Soochow University, Suzhou Industrial Park 215123, Jiangsu, China
b
UMR 1253 STLO, Agrocampus Ouest, INRA, F-35042 Rennes, France
c
Lactalis Recherche et D
eveloppement, F-35240 Retiers, France

a r t i c l e i n f o a b s t r a c t

Article history: Background: Probiotic and starter bacteria are generally dried to produce easy-to-use ingredients that
Received 1 August 2016 are stable and flexible for applications in the food, feed and pharmaceutical industry. The overall demand
Received in revised form for dried probiotic bacteria has increased in the context of a rapidly growing market, evidencing the need
6 February 2017
for their larger scale production.
Accepted 7 February 2017
Available online 6 March 2017
Scope and approach: The spray-drying of bacteria enables a larger production scale than the freeze-
drying currently used; energy costs are lower and the process is sustainable. This is also a promising
way to microencapsulate bacteria within various protective matrices to ensure their improved resistance
Keywords:
Spray drying
during storage, technological processes and digestive stresses.
Probiotics Key findings and conclusions: This review highlights some key strategies to improve the viability and
Stress tolerance efficacy of probiotics spray-drying, such as the enhancement of bacterial resistance, improved protection
Storage of the drying medium and optimization of the drying process. It also focuses on factors during the pre-
Viability and post-drying stages which may influence the quality and efficacy of spray-dried probiotic powders.
Functionality © 2017 Elsevier Ltd. All rights reserved.

1. Introduction preserve probiotic bacteria; but it is a time-consuming and


expensive process (Table 1). Among the drying techniques that are
In recent years, increasing numbers of bacteria have been possible, spray-drying is one of the most predominant in the dairy
investigated for their probiotic potential, as they confer health industry (Schuck et al., 2016). It consists in spraying the liquid feed
benefits on the host when administered live and in appropriate in fine droplets (10e150 mm) that are directed into a flow of hot and
quantities. The beneficial effects of probiotics depend on the spe- dry air (usually 150  Ce250  C). The increase in the air-liquid
cific strain or specie, the dose and viability of the bacteria ingested interface area subsequent to spraying dramatically increases the
(Hill et al., 2014). The International Dairy Federation (IDF) recom- drying kinetics, and it is generally admitted that drying occurs
mends a minimum of 107 live probiotic bacterial cells per gram or within a few seconds. When compared to freeze-drying, spray-
milliliter of product at the time of consumption (Corona-Hernandez drying represents a lower specific energy cost and higher produc-
et al., 2013). Therefore, maintaining adequate levels of viable cells tivity (Table 1). There remain challenges associated with the use of
and ensuring their properties throughout shelf-life is a prerequisite spray-drying to produce viable cultures, especially with “sensitive”
for their further use, e.g. when incorporated in food products and probiotic strains (Broeckx, Vandenheuvel, Claes, Lebeer, & Kiekens,
during the digestion process. 2016; Fu & Chen, 2011; Peighambardoust, Golshan Tafti, & Hesari,
Drying is a widely-used process for food preservation, ensuring 2011).
a stable and extended shelf-life, reducing transportation costs and This review offers an update of the state-of-the-art on the
facilitating trade. Freeze-drying remains the preferred technique to adaptive response of probiotic bacteria to several stresses related to
spray-drying conditions (Fig. 1). We also review recent advances in
the preparation and spray-drying conditions for probiotic culture
that have been shown to be protective or have a positive impact on
* Corresponding author. UMR1253 STLO, 65 rue de Saint-Brieuc, F-35042 Rennes, probiotic viability.
France.
E-mail address: romain.jeantet@agrocampus-ouest.fr (R. Jeantet).

http://dx.doi.org/10.1016/j.tifs.2017.02.007
0924-2244/© 2017 Elsevier Ltd. All rights reserved.
2 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

Table 1
, and Me
General production pattern and specific energy consumption of commonly used concentration and drying processes. Data from Bimbenet, Schuck, Roignant, Brule jean
(2002) and Schuck et al. (2015).

Processes Production pattern Specific energy Productivity Advantages Limits


consumption (ton.year1)
(kJ.kg1 water)

Membrane Liquid / Concentrate Batch 40 n.a. Low temperature and energy cost Fouling, non-specificity of the cut-off,
separation Continuous cost of the replacement of membranes
a
Spray- drying Liquid / Solid Continuous 5300 ~50000 Low heat treatment and residence time High investments
Freeze-drying Liquid / Solid Batch 18,000 ~10000a Very low heat treatment, production High residence time (24e48 h)
of porosity, improvement of rehydration
Fluidized-bed Solid / Solid Continuous 11,400 n.a. Low heat treatment, production Very high specific energy consumption,
drying of powder below Tg high residence time and investments

n.a.: not available.


a
An approximate value for a large scale dryer from personal communication (The productivity actually depends on equipment, production scale and market demand).

Fig. 1. The most influential factors and adverse stresses experienced during the growth, spray-drying, storage and application of probiotics.

2. Pre-drying stage obtained, with around 100% survival and 1010 CFU g1 cell counts in
the powder (Schuck, Dolivet, Me jean, Herve , & Jeantet, 2013). In our
2.1. Selection of bacterial strains recent work, Propionibacteria freudenreichii ITG P20 was found to
survive better than Lactobacillus casei BL23 (70% versus 40%), even
To date, most efforts have focused on the drying of Lactobacillus, under harsher drying conditions (180  C Tinlet, 73  C Toutlet versus
Lactococcus and various Bifidobacteria species. These probiotic 140  C Tinlet, 63  C Toutlet) (Huang et al., 2016a). However, compared
bacteria generally do not survive well after spray-drying because of to Lactobacillus (Lb.) and Bifidobacteria, studies involving the drying
the harsh conditions which prevail during the process (Table 2). The of Propionibacteria (P.) are still rare.
resistance characteristics of a bacterial strain should thus constitute Streptococcus (S.) is usually more resistant than Lactobacillus to
an important criterion when selecting probiotic bacteria, in order to spray-drying; for instance, S. thermophilus was shown to survive
improve the final probiotic viability of the spray-dried powders. better than Lb. delbrueckii ssp. bulgaricus in spray-dried yoghurt
Heat, osmotic, oxidative and desiccation stresses are usually (Bielecka & Majkowska, 2000; Kumar & Mishra, 2004), and
considered to be the main mechanisms which cause the inactiva- S. thermophilus CCRC14085 survived better than Lb. acidophilus
tion of bacteria during and after spray-drying (Santivarangkna, CCRC 14079 in spray-dried fermented soymilk (Wang, Yu, & Chou,
Kulozik, & Foerst, 2008). It has been shown that different bacte- 2004). The threshold temperature at which damage is caused to
rial species, or even strains, may display variable tolerance towards microbial cells is usually within the range of the upper limit of
such stresses. growth temperature of the microbial species (Foerst & Kulozik,
By comparison with Lactobacillus, Lactococcus and Bifidobacteria, 2011). Thus the spray-drying resistance of S. thermophilus is prob-
Propionibacteria, whose probiotic properties were reviewed by ably linked to its greater thermotolerance. In another observation,
Cousin, Mater, Foligne , and Jan (2010), usually display higher Lb. paracasei NFBC 338 was however found to survive as success-
tolerance due to their greater abilities for environmental adapta- fully as S. thermophilus (Kearney et al., 2009). This finding indicates
tion, either through their metabolism or a multi-tolerance response that certain strains within a usually fragile specie may be as resis-
(Huang et al., 2016b; Leverrier, Vissers, Rouault, Boyaval, & Jan, tant as bacteria from a generally robust specie.
2004). Using sweet whey (50% w/w) as the drying medium, two When compared within the Lactobacillus genus, Lb. plantarum is
Propionibacteria acidipropionici strains were spray dried with a pilot a specie with relatively robust stress tolerance (Ferrando,
spray dryer under around 140  C inlet temperature and 60  C outlet Quiberoni, Reinhemer, & Sua rez, 2015). Mille, Beney, and Gervais
temperature. A high degree of viability following spray-drying was (2005) showed that the osmotic tolerance of Lb. plantarum was
Table 2
Review of key factors determining viability of probiotic bacteria (Lb. refers to Lactobacillus, Lc. refers to lactococcus, S. refers to Streptococcus, B. refers to bifidobacteria, P. refers to propionibacteria) upon spray-drying.

Bacteria Growth parameters Cell concentration Drying parameters Devices Powder Survival Reference
before drying moisture (%)
Medium T ( C) Medium Tinlet ( C) Toutlet ( C)
(CFU mL1) content (%)

S. thermophilus MK-10 Condensed non fat 42 9.5  108 Fermented growth medium n.a.g 60 “Luwa” type industrial 10.2 69.5 (Bielecka &
milk (18% w/w) (18% w/w) 80 spray dryer 4.4 12.7 Majkowska, 2000)
Lb. bulgaricus 151 60 10.2 22.1
80 4.4 8.0
Lb. paracasei NFBC338 RSMh (20% w/w) with yeast 37 ~108 Fermented growth medium 170 85e90 Buchi B-191 Mini dryer, ~3.5j ~50j (Gardiner et al., 2000)
Lb. salivarius UCC118 extract (0.5% w/w), (adjust pH to 6.8) Switzerland ~1j
additional sucrose (1% w/w)
for strain UUC118
l
B. infantis CCRC14633 MRS broth with 0.05% cysteine 37 n.a. Gelatin (10%) 100 50 Buchi B-191 Mini dryer, 10.0 1.3 (Lian et al., 2002)
B. infantis CCRC14661 Switzerland 0.2
B. longum ATCC15708 54.3
B. longum CCRC14634 39.4
B. longum B6 7.8 63.7

S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17


Lb. rhamnosus E800 MRS broth 37 3.2  109j RSM (20% w/v) 170 85e90 Buchi B-191 Mini dryer, 3.8 ~30j (Corcoran et al., 2004)
Lb. rhamnosus GG 1.6  109j Switzerland 3.7 ~50j
Lb. salivarius UCC500 6.3  109j 2.7 ~1j
17 strains of MRS broth with 0.05% cysteinel 37 0.1e3.2  109 RSM (20% w/v) 170 85e90 Buchi B-191 Mini dryer, 2.5e4.2 12~102a (Simpson et al., 2005)
Bifidobacterium species Switzerland
B. animalis ssp. lactis BB12b 5.0  108
3.2 79
B. breve NCMB8807b 6.3  108 3.2 38
Lb. rhamnosus GG MRS broth supplemented 37 1.8  109 Trehalose (20% w/w) n.a. 65e70 Buchi B-191 Mini dryer, 3.8 69 (Sunny-Roberts &
l
with trehalose (1.25% w/v) Trehalose (20% w/w) þ Switzerland 80.8 Knorr, 2009)
Monosodium glutamate
Lb. rhamnosus E800 Trehalose (20% w/w) 4.1 23
Trehalose (20% w/w) þ 89.3
Monosodium glutamate
Lb. paracasei A12 MRS broth 37 4.0  108 RSM (20% w/v) 170 85 Buchi B-290 mini spray <4 ~100j ez et al., 2013)k
(Pa
Lb. casei Nad 4.0  108 drier, Switzerland
Lb. acidophilus A9 5.0  109
P. acidipropionici Acid whey permeate (10% w/v) 30 1010e Sweet whey (~50% w/w) 130 60 Niro Atomizer (GEA, ~3d 100 (Schuck et al., 2013)
with corn steep (2.5% w/v) and Germany) pilot Bionov
yeast extract (1% w/v) spray dryer, France
c
14 strains of Lc. lactis species M17 broth with 0.5% w/v glucose 30 n.a. RSM (20% w/v) 200 100 Buchi B-290 mini spray <4 (Dijkstra et al., 2014)
drier, Switzerland
Lb. acidophilus La-5 MRS broth 37 ~109j RSM (30% w/w) 180 85e95 SD-05 lab-scale spray- 4.3 77.7 (Maciel et al., 2014)
Sweet whey (30% w/w) dryer, Lab-Plant, UK 4.8 75.4
Lb. plantarum WCFS1 MRS broth 30 ~108 Maltodextrin (20% w/w) 135 90 Buchi B-290 mini spray 4e8 ~10j (Perdana et al., 2014)
drier, Switzerland
j
Lb. plantarum 299v MRS broth 37 n.a. Orange juice (1.5% w/v) þ 150 70 Niro spray dryer, aw ¼ 0.42 ~100 (Barbosa et al., 2015a)
Maltodextrin (2% w/v) Denmark
Lb. casei 431 MRS broth 37 5.0  108 Whey protein isolate and gum 180 80 Buchi B-290 mini spray 2.90 37.6 (Eratte et al., 2015)
arabic (3:1 mixture, ~5% w/w) drier, Switzerland
2.0  109 Whey protein isolate, tuna 3.19 56.2
oil and gum arabic (3:2:1
mixture, ~10% w/w)
Lb. plantarum CNRZ 1997 MRS broth 37 ~108 Phosphate-buffered 145 70 Buchi B-290 mini spray aw ¼ 0.3 ~50j (Iaconelli et al., 2015)
Lb. zeae CNRZ 2268 saline (~1% w/v) drier, Switzerland ~1j
l
B. bifidum CIP 56.7 MRS broth with 0.05% cysteine ~10j
Lb. zeae LB1 MRS broth 37 ~1010 Sodium caseinate (10% w/w) 170 80 Laboratory-scale ADL 6.80 ~15 (Liu et al., 2015)
Sodium caseinate and vegetable 310 spray dryer, USA 3.25 ~15
oil (1:1 mixture, 20% w/w)
(continued on next page)

3
Table 2 (continued )

4
Bacteria Growth parameters Cell concentration Drying parameters Devices Powder Survival Reference

before drying  
moisture (%)
Medium T ( C) Medium Tinlet ( C) Toutlet ( C)
(CFU mL1) content (%)

Sodium caseinate and low 3.68 ~65


melting point fat (1:1 mixture,
20% w/w)
Lb. acidophilus LA-5 MRS broth 37 3.2  108j Goat's milk (20% w/v) 195 85 Buchi B-290 mini spray 2.9 ~1j (Ranadheera, Evans,
B. animalis subsp. lactis BB-12 RC medium l 108j drier, Switzerland < 1j Adams, & Baines, 2015)
P. jensenii 702 SL broth 30 109j ~8j
Lb. casei BL23 Sweet whey (30% w/w) with 37 2.0  109 Fermented growth medium 140 63 Niro A/S spray dryer, 6 40 (Huang et al., 2016a)
P. freudenreichii ITG P20 casei peptone (0.5% w/w) 30 2.5  109 180 73 Denmark 70
Lb. plantarum A17 MRS broth 37 3e4  109 i
WPI (pH 7.0, 10% w/w) 110 68e70 Laboratory LabPlant 5.6 69.0 (Khem et al., 2016)
WPI (pH 4.0, 10% w/w) SDBasic FT30MKIII, 5.3 39.3
Denatured WPI (pH 7.0, 78  C spray dryer, UK 5.4 25.0
for 20 min, 10% w/w)
l
B. infantis ATCC 15679 MRS broth with 0.05% cysteine 37 108~109 Maltodextrin (15% w/v) 110 75e80 Laboratory LabPlant aw ~ 0.24 91.6 (Bustamante et al., 2017)
Lb. plantarum ATCC 8014 MRS broth Maltodextrin (15% w/v) SD-05 spray dryer, 83.0

S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17


Maltodextrin, Chia seed England 98.0
mucilage and Chia
seed soluble protein
(7.5:0.6:7.5 mixture, ~15% w/v)
Lb. rhamnosus GG MRS (pH controlled 37 ~3.2  107e Micellar caseins with denatured 195 85 Anhydro MicraSpray 6.3 ~50j (Guerin et al., 2017)
at 6.8 by 6 M NaOH) whey proteins (90:10 v/v 150, Danmark
mixture, 12.5% w/w)
~4  108e Chymosin treated micellar caseins 5.8 ~50j
with denatured whey proteins
(90:10 v/v mixture,
12.5% w/w)
Lb. casei BL23 Sweet whey (30% w/w) 37 1.6  109 Fermented growth medium 140 60 Niro Atomizer (GEA, 5.2 ~60 (Huang et al., 2017)
127 47 Germany) pilot Bionov 5.5f ~100f
spray dryer, France
P. freudenreichii ITG P20 30 5.0  109 140 60 5.2 ~100
127 47 6.0f ~100f
a
Survival ranged from 12% to 102%, depend on species and strain, closely related strains exhibiting superior heat and oxygen tolerance performed best after spray drying.
b
Representitive examples from the 17 strains.
c
Lc. lactis subsp. lactis strains generally displayed more robust phenotypes than L. lactis subsp. cremoris strains. The most robust strains displayed a more-than-200-fold-better survival during spray drying than the most
sensitive strains.
d
Data was calculated from total solid content of spray dried powders after delivery by a crystallizer.
e
The unit of bacterial population is CFU g1.
f
Data was from final powders at the end of multi-stage drying.
g
n.a.: not available.
h
RSM: Reconstituted skim milk.
i
WPI: Whey protein Isolate.
j
Means data were calculated from the figures or information from the paper.
k
In vivo study of effect of spray drying on probiotic functionality.
l
Growth under anaerobic condition.
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 5

generally greater than that of Escherichia coli, Bradyrhizobium susceptibility to both process and environment is strain-dependent,
japonicum and Lb. bulgaricus, and in some cases even higher than drying conditions need to be adapted while considering the latter
that of yeasts (Saccharomyces cerevisiae and Candida utilis). Other and not directly transposed from one strain to another.
species frequently studied in the context of spray-drying include Lb.
rhamnosus, Lb. casei, Lb. paracasei, Lb. salivarius and Lb. acidophilus. 2.2. Response to cellular stress
However, the stress tolerance of these species appears to be strain-
dependent (Table 2). Growth conditions are widely accepted as key determinants for
Bifidobacterium longum ssp. longum B6 survived spray-drying the acquisition of bacterial stress tolerance. Indeed, like other
better than Bifidobacterium longum ssp. infantis CCRC 14633 when bacteria, probiotics are able to withstand and adapt to different
fermented soymilk and reconstituted skim milk were used as the adverse environmental stresses (e.g. high temperatures or osmotic
drying media, while both B. longum. ATCC15708 and CCRC 14634 stress) by activating the cellular stress-response system, which
survived better than B. infantis CCRC14661 when using gelatin as determines their tolerance during spray drying. This system is
the spray-drying medium (Lian, Hsiao, & Chou, 2002; Wang et al., generally induced by exposure to a sub-lethal dose of the stress
2004). In a comparison of 17 strains from the Bifidobacterium itself (Table 4). For instance, significant improvements in heat and
genus, it was found that the greater the resistance to heat and spray-drying tolerance have also been reported for heat-adapted
oxygen, the better was survival after spray-drying (Simpson, Lb. salivarius (Zhang, Lin, & Zhong, 2016).
Stanton, Fitzgerald, & Ross, 2005). This relationship has also been In addition to heat adaptation, the exposure of bacteria to sub-
reported in the Lb. rhamnosus and Lactococcus lactis (Lc. Lactis) lethal osmotic stress may also confer tolerance to heat and spray-
species (Dijkstra et al., 2014; Lavari et al., 2015). drying. For example, in our recent work, the survival of both Lb.
Although the species may generally reflect the apparent resis- casei BL23 and P. freudenreichii ITG P20 after spray-drying was also
tance of different bacteria, the robustness of probiotic bacteria improved significantly after harsh hyperosmotic conditions during
during spray-drying seems more likely to be strain-dependent. their growth in concentrated sweet whey (Huang et al., 2016a,
Such variability is likely related to the environment of the original 2016b). Although the improvement in spray-drying survival was
source, the presence of specific genes, interactions with the extra- not as effective as for heat or salt adaptation, the exposure of
cellular matrix, or an ability for intracellular polyphosphate accu- bacteria to H2O2 (0.003 M for 30 min) or bile salts (0.1% w/v for
mulation and exopolysaccharide production, etc (Fig. 2 and Table 3). 30 min) could also improve bacterial viability during spray-drying
(Lebeer, Vanderleyden, & De Keersmaecker, 2008). Given that (Desmond, Stanton, Fitzgerald, Collins, & Ross, 2001).

Fig. 2. The intracellular accumulation of polyphosphate by P. freudenreichii was visualized by DAPI staining prior to observation under a confocal microscope. P. freudenreichii culture
in (A) YEL broth and (B) sweet whey medium (30% w/w). Green fluorescence indicates cytosolic polyphosphate, and blue fluorescence DNA. The culture in the sweet whey medium
displayed higher tolerance against heat, acid, bile salt and spray-drying stress, as well as improved stability during storage. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

Table 3
Examples of compatible solute accumulation on improvement of bacterial stress tolerance.

Compatible solute Bacterial strain Stress response Application in spray-drying Reference

Proline Lb. acidophilus IFO 3532 Enhance osmotic tolerance not to date (Jewell & Kashket, 1991)
Carnitine Lb. plantarum P743 Enhance osmotic tolerance not to date (Kets & Bont, 1997)
Acetylcarnitine
Propionylcarnitine
Glutamate Lb sakei CTC 494 No significant effect on heat tolerance Improved the survival (Ferreira et al., 2005)
Glycine betaine Pantoea agglomerans CPA-2 Enhance osmotic, water and heat tolerance not to date  et al., 2005)
(Teixido
Ectoine
Listerial betaine Lb. salivarius UCC118 Enhance osmo-, cryo-, baro- and chill tolerance Improved the survival (Sheehan et al., 2006)
Glycogen P. freudenreichii ITG 20 Enhance osmotic, heat, acid and bile-salt tolerances Improved the survival (Huang et al., 2016b)
Trehalose
Polyphosphate
6 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

Table 4
Examples of genes and proteins involved in probiotics adaptation towards spray-drying-related stress.

Stressa Gene/Protein Description and function Example strain Reference

General rpoS/s38 or sS Transcription initiation factor that enables specific binding of E. coli Nissle 1917 (Coldewey et al., 2007)
RNA polymerase to gene promoters at stationary phase
Heat Osmotic dnaK, dnaJ, grpE/DnaK, Heat shock proteins folding chaperones to prevent misfolding Lb. casei Zhang (Wu et al., 2011)
DnaJ, GrpE and to promote the refolding or degradation of unfolded
groL, groS/GroEL,GroES polypeptides Lb. paracasei NFBC 338 (Corcoran et al., 2006)
hsp20/Hsp20 P. freudenreichii ITG P20 (Huang et al., 2016b)
clp genes/ClpB, ClpC, Clp ATPase family members act as chaperones and regulators of P. freudenreichii ITG P20 (Huang et al., 2016b)
ClpE, ClpP etc. proteolysis
htrA/HtrA High temperature requirement protein with dual chaperone- Lb. plantarum FS5-5 (Wu et al., 2016)
protease activities
Heat ftsH/FtsH ATP-dependent zinc metalloprotease with dual chaperone- Lb. plantarum WCFS1 (Bove et al., 2012)
protease activities and involving in protein quality control
Osmotic opu genes/OpuA, ATP-binding ABC transporter in charge of compatible solute P. freudenreichii ITG P20 (Huang et al., 2016b)
OpuB, OpuC etc. transport
ptsH/HPr Phosphocarrier protein family catalyzes the phosphorylation of
incoming sugar substrates concomitantly with their
translocation across the cell membrane.
Oxidative ahpC/AhpC Alkyl hydroperoxide reductase C22 protein destroys toxic B. longum NCC2705 (Zuo et al., 2014)
radicals normally produced within cells to maintain cell redox
homeostasis
katA/KatA Catalase decomposes hydrogen peroxide into water and Lb. brevis CGMCC1306 (Lyu et al., 2016)
oxygen, to maintain cell redox homeostasis
nox,npr/NOX, NPR NADH oxidase and NADH peroxidase promote NADH cycling to Lb. plantarum C17 (Zotta, Guidone,
maintain cell redox homeostasis Ianniello, Parente,
& Ricciardi, 2013)
sodA/SodA Superoxide dismutase destroys toxic radicals normally Lb. casei BL23 (Lee, Tachon, 2015)
produced within cells to maintain cell redox homeostasis
a
Cross-protection possibly emerges between different stresses.

Ananta and Knorr (2003) also found that Lb. rhamnosus GG pre- Herman, White, and Vesey (2006), this result suggests that either
treated at high pressure (100 MPa) displayed greater resistance to bacteria do not need to synthesize new proteins during recovery
heat. Furthermore, starvation (mostly glucose starvation) can also from drying injury, or that these injured bacteria are unable to
induce bacterial tolerance of osmotic and heat stresses (Guchte synthesize proteins. In this case, the overproduction of stress
et al., 2002). response proteins before drying may constitute an effective means
Apart from the adaptation to environmental conditions, modest of protecting bacteria from drying injury, and particularly those
improvements to stress tolerance can also be achieved through proteins or chaperones responsible for preventing or repairing
genetic engineering. For instance, a genetically engineered Lb. misfolding polypeptides.
paracasei strain was able to overproduce the heat shock protein
GroESL, and thereby exhibited a 10-fold higher thermotolerance
2.3. Growth conditions and preparation of feed concentrate
than its parent wild-type strain (Desmond, Fitzgerald, Stanton, &
Ross, 2004). Given its heat robustness, the GroESL-overproducing
2.3.1. Growth media
Lb. paracasei strain survived 10-fold better than the wild-type
Only a few published studies have addressed how the growth
strain during spray drying (Corcoran, Ross, Fitzgerald, Dockery, &
medium affects the viability of probiotics after drying. Most studies
Stanton, 2006). Cells appear to be more robust and intact in
have focused on the media used in laboratories, such as De Man
GroESL-overproducing bacteria and are sometimes huddled
Rogosa and Sharpe medium (MRS broth) which is non-food grade,
together. However, the overexpression of GroESL did not enhance
and relatively expensive for industrial application. As an example,
survival during storage of the lactic acid bacteria (LAB) powder.
the presence of 1 mol L1 or 1.25 mol L1 NaCl in MRS during the
Moreover, the thermotolerance of this GroESL overproducing strain
growth of Lb plantarum resulted in decreased residual activity after
can be further improved up to 50-fold after sub-lethal heat adap-
drying in a fluidized bed or by convection. The viability of Lb.
tation, which indicates the involvement of other heat stress
plantarum after drying was also greater when the bacteria were
response mechanisms (e.g., contributions from other heat shock
grown in diluted MRS medium, despite lower betaine and carnitine
proteins or general stress proteins) (De Angelis, Calasso, Cavallo, Di
uptake in this case (Linders, Meerdink, & Van't Riet, 1997).
Cagno, & Gobbetti, 2016; Lebeer et al., 2008). Hence, the induction
As well as these studies based on MRS medium, several others
of the whole battery of heat stress tolerance mechanisms by sub-
have explored the effects of other media on growth and the sub-
lethal heat treatment may procure greater viability.
sequent drying or storage of bacteria. Dairy media, such as skim-
A genetically engineered Lb. salivarius UCC 118, which can
med milk and whey, were mostly used during these studies
overexpress the listerial betaine uptake system BetL, was also re-
(Table 2). By comparison with MRS, these media are inexpensive,
ported to have a 5-fold improvement in survival following spray-
easily used for mass production and edible. Moreover, they may
drying (Sheehan, Sleator, Fitzgerald, & Hill, 2006). Indeed, betaine
improve following production processes such as drying (Huang
protects lactic acid bacteria (LAB) from desiccation (Table 3).
et al., 2016a; Lavari, P
aez, Cuatrin, Reinheimer, & Vinderola, 2014).
To sum up, in most cases, bacterial spray-drying tolerance may
When facing a reduction in water activity, bacteria tend to
be affected by how the strain or culture has been produced.
compensate for the concomitant osmotic pressure and thus main-
Furthermore, Teixeira, Castro, Malcata, and Kirby (1995) reported
tain their viability by accumulating compatible solutes such as
that protein synthesis was not observed in Lb. bulgaricus during its
amino acids, quaternary amines or carbohydrates, etc (Table 3).
recovery from spray-drying injury. As concluded by Morgan,
(Kets, Teunissen, & Debont, 1996; Wood, 2011). The accumulation
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 7

of compatible solutes from media during growth has thus been keep the growth medium and directly spray-dry the probiotic
used as a strategy to increase the viability of probiotics during culture during large-scale production. Apart from drying probiotic
drying (Huang et al., 2016a). However, it has been suggested that cells, some beneficial metabolites such as short chain fatty acids,
the uptake of compatible solutes may not necessarily guarantee vitamins and bacteriocin are also retained, possibly improving the
tolerance towards drying, depending on the bacterial strains use value of powders. This process requires the dual use of one
studied (Morgan et al., 2006). Furthermore, it is worth noting that medium for both bacterial growth and spray-drying. A food grade
the synthesis of compatible solutes in LAB was seen to be depen- medium is highly desirable because of application considerations.
dent on the presence of their precursors in the medium (Wood, However, a key problem with this process is that the solid content
2011). The transport and accumulation of these compatible sol- of bacterial growth isotonic media is typically low (5%e10%), while
utes is also an energy-dependent process (Romantsov, Guan, & a high solid content in the feed concentrate is required to enable
Wood, 2009). Because of the short timescale of the spray-drying productivity and energy saving. This is usually achieved by adding
process, it has been suggested that compatible solutes or their more powders to the culture, or by implementing concentration
precursors should be added during the growth of bacteria rather techniques (Schuck et al., 2013). However, these intermediate steps
than before spray-drying (Huang et al., 2016a). may cause a loss of viability and increase the risk of contamination.
Our recent work revealed that it is possible to grow Lb. casei and
2.3.2. pH of growth P. freudenreichii in highly concentrated sweet whey with a total
The effects of the pH of a medium on bacterial viability during solid content of up to 30% (Huang et al., 2016a, 2016b). Interestingly,
drying remain a matter of discussion, depending on the processing the final biomass yield was improved when compared to that of
scheme implemented. Linders et al. (1997) reported a two-fold growth in isotonic 5% sweet whey, along with the consequent
increase in Lb. plantarum viability after drying when the pH was higher bacterial resistance against spray-drying.
controlled during growth. By contrast, the viability of Lb. bulgaricus
during spray-drying and heating was found to be greater when the 3. During spray-drying
pH was not controlled during growth (Silva et al., 2005). It was
shown that in this case, enhanced protection was due to the over- 3.1. Bacteria may be injured during spray-drying
expression of Hsp70, GroES and GroEL, induced by acid stress (Silva
et al., 2005). However, cells may lose their viability and activity Damage to bacteria during spray-drying is not only ascribed to
when they are unable to maintain a near neutral intracellular pH in the thermal effect, but also to a loss of bound water at the cell
a low pH environment for a long period. The ability to maintain surface. One of the sites most susceptible to cellular injury is the
intracellular pH homeostasis differs between strains (Baker-Austin cytoplasmic membrane, and this is exacerbated during spray-
& Dopson, 2007). This may condition strain dependence relative to drying (Santivarangkna et al., 2008). Indeed, the removal of water
acid adaptation and thus induced cross-protection against drying. leads to a state transition of the phospholipid bilayer, from lamellar
Further, because acid stress influences cells in a dynamic manner, to gel phase, or even a hexagonal one. This results in the phos-
the final pH of cultures, the time of harvest or feed storage time pholipid chains gaining a rigid and fully extended structure (Crowe,
before drying should all be taken into account when investigating Carpenter, & Crowe, 1998; Leslie, Israeli, Lighthart, Crowe, & Crowe,
the effects of growth pH. 1995). It was indeed shown that fine holes at the surface of Lc.
cremoris were observed by scanning electron microscopy in single
2.3.3. Growth phase droplet drying with skimmed milk (10% w/w) at 90  C for 5 min and
Probiotics are generally harvested during either the late expo- 110  C for 4 min, while the cell surface remained intact when drying
nential phase or the early stationary phase, at which a maximal at 70  C for 6 min (Fu, Woo, Selomulya, & Chen, 2013).
yield is attained. However, (early) stationary phase is mentioned Other dehydration stress targets include the nucleic acids
more frequently as being optimal and used than the (late) expo- (where the mechanism remains unclear) and ribosomes, which are
nential phase, because the cells collected also exhibit higher probably injured as a result of the escape of Mg2þ from the heat-
viability during drying (Peighambardoust et al., 2011). The viability compromised cell membrane (Nierhaus, 2014).
of Lb. rhamnosus was 2%, but over 14% and 50% when cells were
harvested in the lag, early exponential and stationary phases, 3.2. Drying devices
respectively (Corcoran, Ross, Fitzgerald, & Stanton, 2004). Indeed,
resistance to various types of stress is increased during the sta- 3.2.1. Scale of spray-drying
tionary phase, as its challenging conditions trigger a stringent To date, little attention has been paid to the effects of spray-
response that leads to multi-tolerance which requires general drying devices on the viability of probiotics in powders. The main
stress proteins and alternative sigma factor (Alcantara, Revilla- influence of different devices on probiotic powders is probably the
Guarinos, & Zuniga, 2011; Hussain, Knight, & Britz, 2009; Upa- residence time of particles in the drying chamber: The longer the
drasta, Stanton, Hill, Fitzgerald, & Ross, 2011). residence time, the longer the bacteria are exposed to stress and
consequently the poorer the viability. Another factor worth noting
2.3.4. Harvesting techniques is that industrial scale spray dryers are normally suitably equipped
Centrifugation is still the most widely used harvesting tech- with pneumatic devices to enable the continuous collection and
nique. It has long been considered to be efficient in concentrating cooling of the powders, thus maximizing viability.
cells. The usual harvesting temperature is 4  C. However, the effects Table 2 presents the drying equipment used during different
of these centrifugal conditions on bacterial viability after drying studies, which mainly focused on laboratory-scale spray dryers. It is
have rarely been reported. As well as centrifugation, other har- worth noting that only three studies mentioned experiments on a
vesting techniques, such as membrane filtration, were mentioned larger scale with a water evaporation capacity in the range
in the review by Santivarangkna, Kulozik, and Foerst (2007), but to 100e1645 kg h1) (Bielecka & Majkowska, 2000; Huang et al., 2017;
the best of our knowledge, they have not been reported in the Schuck et al., 2013). Although recent work has shown that there
context of the spray-drying of bacteria. were no significant differences in bacterial survival after spray-
Separating bacteria from growth medium can cause consider- drying at a laboratory scale (~5 kg h1) and in a large pilot-scale
able wastage of materials and energy. It is therefore preferable to dryer, it is still difficult to conclude as to the influence of scaling-
8 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

up on the residual probiotic viability of the resulting powders reducing disaccharides trehalose and sucrose, the effect of the
(Huang et al., 2017). reducing disaccharide lactose in protecting bacterial cells during
drying remains doubtful, because of the possibility of its interaction
3.2.2. Atomization with proteins (Maillard reaction) during drying and storage. When
The cell inactivation caused by the atomization process per se is using whey permeate (which mainly consisted of lactose) as a
usually silent or suggested to be negligible. Fu, Suen, and Etzel carrier for the spray-drying of Lb. acidophilus, its survival was only
(1995) reported a complete viability of Lc. lactis ssp. lactis after half of that achieved with RSM under similar conditions (Riveros
the spraying step, despite the shear force experienced in the nozzle. et al., 2009). Nevertheless, whether or not the lactose in RSM
By contrast, Riveros, Ferrer, and Bo rquez (2009) found that when plays the main role in stabilizing LAB during storage remains
lowering the two-fluid spray nozzle pressure from 1.5 bar to 1 bar, undetermined.
the viability of Lb. acidophilus increased by around 2 log CFU g1. The proteins in RSM can prevent cellular injury by stabilizing
However, it should be noted that a lower spray pressure might first cell membrane constituents. Furthermore, they may form a pro-
of all affect the droplet size and hence the drying kinetics and tective coating on the bacterial cell wall, when interacting with
subsequent temperature increase. Besides, Guerin et al. (2017) milk calcium (Huang et al., 2014; Zheng, Fu, Huang, Jeantet, & Chen,
showed that the increase in bacteria population after atomization 2016). The calcium in milk can also cause milk protein aggregation
might be caused by cell reorganization during atomization, or more during heat treatment, and indeed the protection of Lb. rhamnosus
specifically, the Lb. rhamnosus GG distributed in linear chains before GG was higher than that of untreated milk during thermal
atomization and in individual cells after atomization. During convective droplet drying (Wang, Huang, Fu, Jeantet, & Chen, 2016).
another study, a significant reduction in survival was seen for a This improved protection has been hypothesized to be linked to the
strain of Lc. lactis ssp. cremoris after atomization using both a rotary gel-like structure of calcium-aggregated milk, which displays rapid
wheel atomizer and two-fluid nozzle (Ghandi, Powell, Howes, water-like drying kinetics and forms a porous particle. Interest-
Chen, & Adhikari, 2012). Furthermore, the addition of ascorbic ingly, innovative spray-dried microparticles with different recon-
acid (as an anti-oxidant) to the drying medium was found to reduce stitution behaviors have also been developed based on milk protein
the damage caused by atomization. These results indicate that a aggregation controlled by chymosin (Guerin et al., 2017). The mi-
loss of cell viability during atomization may be strain-dependent, croparticles displayed an encapsulation effect on Lb. rhamnosus GG
and the inactivation mechanism could be triggered by both shear when rehydrating at 40  C, while a releasing effect was seen at 8  C.
force and oxidative stress. During another laboratory-scale spray-drying experiment, RSM
(20% w/v) was found to confer superior protection on Lb. paracasei
3.3. Protective agents Nad, Lb. casei A13 and Lb. acidophilus A9, with around 100% survival
for all three strains under the same drying conditions (Pa ez et al.,
A strategy that is commonly employed to enhance cell protec- 2013). As well as this well-preserved viability, the resistance to
tion during drying consists in adding specific components, known digestion and immunomodulation capacity of these three strains
for their protective properties, to the surrounding medium. within the powders were also improved significantly when
compared to the fresh culture.
3.3.1. Carbohydrates
Among protective agents of a carbohydrate type, trehalose has 3.3.3. Other protective agents
by far been the one most frequently investigated, given the estab- Other protective agents have also been used in the spray-drying
lished role of its accumulation in the acquired survival of some of probiotics: dairy-based materials have often been reported
microorganisms in a context of anhydrobiosis. It has been sug- because of their compatibility with probiotics and the possibility to
gested that this mechanism is due to the stabilizing effect of promote their efficiency (Lee, Yin, Griffey, & Marco, 2015). Whey
trehalose on membranes and proteins, by replacing the water has also been reported as being an efficient protective agent for the
around polar residues within these macromolecular structures (i.e. spray-drying of probiotics (Huang et al., 2016a, 2017; Maciel,
water replacement hypothesis), and thus decreasing the membrane Chaves, Grosso, & Gigante, 2014). Its advantages are linked to its
phase transition temperature (Crowe, Crowe, & Chapman, 1984; source as a by-product of cheese manufacturing and to its poten-
Morgan et al., 2006). Moreover, Conrad, Miller, Cielenski, and de tially good powder quality (e.g. solubility, flowability, dispersibility,
Pablo (2000) demonstrated a synergistic effect of trehalose and etc.) when compared with casein-based products (Lavari et al.,
borate ions in protecting bacteria: indeed, the glass transition 2014; Sadek et al., 2013).
temperature (Tg) of the dry medium is increased in this case due to Low melting point fat (LMF) has also been reported to protect
crosslinking of the trehalose molecules. Lactobacillus cells from damage when it is added to sodium
In a recent work, carbohydrates that included trehalose, sorbitol, caseinate during spray-drying (Liu et al., 2015). This protection may
mannitol, xylose, glucose, sucrose, maltose, lactose, maltodextrin be attributed to the ability of LMF to absorb some of the thermal
with dextrose, inulin, fructo-oligosaccharides, galacto-oligosac- energy present during spray drying.
charide and potato starch were used as drying matrices in order to As well as dairy-based materials, other reported protective
compare the protection they endowed on probiotic Lb plantarum agents include gelatin, gum arabic, fruit juice, etc. (Table 2), used
WCFS1 during convective droplet drying (Perdana, Fox, Siwei, either alone or in a mixture to achieve synergistic effects. For
Boom, & Schutyser, 2014). It was found that a carbohydrate-rich instance, the spray-drying of probiotics with fruit juice as a me-
formulation with a low molecular weight and high Tg procured dium or additive was recently reviewed by Barbosa and Teixeira
the highest degree of protection during drying. (2016). The incorporation of prebiotics in such protective agents
is also of interest because of the possible “synbiotic” effect. For
3.3.2. Reconstituted skimmed milk example, the addition of Omega-3 fatty acids to a whey protein-
Reconstituted skimmed milk (RSM) appears to be another gum Arabic complex significantly improved the viability of Lb.
suitable medium to enable the efficient spray-drying of probiotic casei 431 after spray-drying (Eratte et al., 2015). When incorpo-
cultures (Table 2). In view of the water replacement hypothesis, the rating galacto-oligosaccharides (GOS) in maltodextrins, or fruc-
lactose in RSM may play an important role (in the same way as tooligosaccharide (FOS) in whey protein isolate (WPI), the viability
trehalose and sucrose) in cell protection. However, unlike the non- of Lb. plantarum after spray-drying was also higher than without
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 9

the addition of GOS or FOS (Rajam & Anandharamakrishnan, 2015; temperatures applied and the expected moisture content and
Sosa et al., 2016). Recently, mucilage and soluble proteins from chia water activity of the dried product.
seed and flaxseed were found to protect Lb. acidophilus, Lb. plan-
tarum and B. infantis during spray-drying (Bustamante, Villarroel, It has thus been shown extensively that the lower the Toutlet, the
Rubilar, & Shene, 2015, 2017). However, Corcoran et al. (2004) higher is post-drying viability (Table 2). Toutlet is therefore consid-
and Pinto et al. (2015) reported that the presence of the pre- ered to be the principal drying parameter that affects the viability
biotics inulin and polydextrose in whey concentrate did not of spray-dried LAB, and any lack of monitoring and control of the
enhance probiotic viability during spray-drying or powder storage. latter may be markedly detrimental (Peighambardoust et al., 2011).
For instance, it was seen that relatively small changes in Toutlet
3.3.4. Total solids content appeared to have significant effects on the survival of Lb. salivarius
The total solids contents of the drying media reported in the (Zhang et al., 2016).
literature are usually 20e30% (w/v): this value has been considered However, apart from bacterial viability, the Toutlet can also in-
as being optimal to ensure the high residual viability of different fluence powder quality: if the Toutlet is too low, this will lead to a
LAB strains (Table 2). An increase in the feed concentration above high residual water activity and moisture content in powders that
this point using gelatin, gum arabic and soluble starch resulted in exceed the values required for prolonged powder storage, e.g.
lower viability of the bifidobacteria (Lian et al., 2002). Indeed, around 0.2 of water activity or 4% of wet basis moisture content
Huang et al. (2016a) explained that the stress adaptation to hyper- (Abe, Miyauchi, Uchijima, Yaeshima, & Iwatsuki, 2009; Vesterlund,
osmolality that drives the increase in viability after drying may not Salminen, & Salminen, 2012). The reduction in viability caused by
be acquired in the case of too short an exposure to osmotic stress an increased Toutlet may also vary with the drying medium used. A
before spray-drying, leading to controversial results on the effect of greater reduction in B. longum and B. infantis viability was observed
the total solid contents. in soluble starch compared to other carriers such as gelatin, gum
Apart from residual bacterial viability, the process energy costs, arabic and skimmed milk (Lian et al., 2002). Further, industrial
powder quality and subsequent product applications should also be spray-drying at a lower Toutlet may also result in improved storage
taken into account. More specifically, a high total solids content is stability (Desmond, Ross, O'Callaghan, Fitzgerald, & Stanton, 2002).
usually desired by industry because of the benefits for drying Therefore, spray-drying temperatures are of considerable signifi-
process productivity, lower energy costs and a better encapsulation cance for the preservation of bacteria, and probably need to be
effect (Fig. 3). By contrast, a low total solids content may make it optimized individually for any new application (Morgan et al.,
possible to maintain an isotonic environment for the bacteria 2006).
before drying and obtain fine powders with a high cell count (a
higher cell/medium ratio) after spray-drying. 3.4.2. Drying rate
The influence of the drying rate on the viability of post-drying
3.4. Drying parameters probiotics is still a matter of debate. At the laboratory scale, it
was found that slow drying kinetics led to significant inactivation of
3.4.1. Drying temperature the dehydration of Lb. plantarum, while a rapid drying rate could
Drying temperature, in combination with drying time, is the key instantly stabilize the cells and thereby prevent this inactivation
factor that influences the final probiotic viability of powders (Perdana et al., 2013). In addition, the influence of a slow drying
because of the general heat sensitivity of these bacterial cells. As rate on the viability of Lc. lactis bv. diacetylactis may have arisen
such, the drying kinetics mainly determines bacterial inactivation from the detrimental reactions favored by higher water mobility
during drying, and particular attention should be paid to the drying above an aW of 0.84e0.88 (Santivarangkna et al., 2007). Conversely,
curve, i.e. temperature as a function of time (Wang et al., 2016). and as already stated in this review, a high drying rate during the
Briefly, the drying curve generally exhibits two stages: first stage of drying, when facilitated by hydraulic membrane
permeability, may limit bacterial adaptation because of exposure
- During the first stage, the temperature of the droplets is limited for too short a time to the gradual withdrawal of moisture (Linders,
to the wet bulb temperature by an almost constant evaporation Meerdink, & Van ’t Riet, 1996).
rate of water, and bacterial inactivation is therefore limited.
- During the second falling rate stage, the temperature of droplets 3.4.3. Process optimization
rises toward the outlet air temperature (Toutlet), depending on In order to obtain probiotic powders with high bacteria viability
the residence time. The latter depends on both the drying and good powder quality, the process should be optimized in order

Fig. 3. (A) Free P. freudenreichii cells, (B) P. freudenreichii cells exposed on the surface of powders spray-dried from a sweet whey (5% w/w) culture, and (C) the surface of powder
spray-dried from a sweet whey (30% w/w) culture. The viability of P. freudenreichii was higher in the 30% sweet whey powders than in the 5% sweet whey powders, during both
spray-drying and storage.
10 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

to simultaneously achieve a reduction in drying temperature and & Patel, 2007). The inactivation kinetics of microorganisms during
an acceptable moisture content in the powders. A multi-stage drying were expressed as:
drying process is the strategy most frequently employed for this
purpose. For example, a process was proposed by Schuck et al.
dðN=N0 Þ
(2013) for pilot-scale spray-drying which enabled a reduction in ¼ kd ðN=N0 Þ (1)
dt
Toutlet during the spray-drying of Propionibacterium acidipropionici,
with sweet whey as the carrier. A crystallizer was used after the where N and N0 are the concentrations of live bacteria (CFU mL1)
spray-drying chamber to enable a low Toutlet at 60  C along with a after and before drying, respectively, and kd is the inactivation rate
high moisture content: indeed, lactose crystallization reduced the constant (s1 for fast drying).
powder hygroscopicity and improved the flowability of the powder. The Arrhenius equation is used to correlate the temperature of
Finally, an external fluid-bed dryer was used to remove the excess the microbial medium with the inactivation rate of the microor-
moisture. This multi-stage drying process was recently improved ganisms. Therefore, kd is usually expressed as:
by replacing the crystallizer with a belt, in order to prevent any
stickiness of the powder (Huang et al., 2017). The Toutlet of the  
E
spray-drying process was further decreased to 47  C, resulting in kd ¼ k0 exp  d (2)
the 100% survival of both Lb. casei BL23 and P. freudenreichii ITG 20. Rg T

3.5. “In-process” investigations where k0 is the pre-exponential factor and Ed is the deactivation
energy which represents the energy required to deactivate living
3.5.1. Single droplet drying cells, and Rg is the universal gas constant (approximately
During a real-life spray-drying process, billions of droplets are 8.314 J mol1K1).
sprayed into a relatively large chamber. It is therefore only possible To incorporate the moisture content factor (X), Meerdink and
to analyze the microbial viability and drying status of samples at Riet (1995) proposed an attractive approach containing just four
the start and end of the process, and not to address the complexity parameters (a, b, k0 and Ed) which must be determined from ex-
of ‘in-process’ droplet-particle conversion and the related mecha- periments. The inactivation rate constant (kd) is then expressed as:
nism of probiotic inactivation. To circumvent these difficulties,
 
single droplet drying has therefore been developed to map the E þ bX
drying behavior of bacteria at the droplet level during the drying
kd ¼ k0 exp aX  d (3)
Rg T
process (Schutyser, Perdana, & Boom, 2012).
Under such a protocol, and controlling the drying conditions, Eq (3) was then further developed by incorporating the influ-
the protective effect of the medium can easily be compared without ence of average drying rate and average heating rate as:
running spray-drying at a larger scale. On the other hand, the effect
   
of certain drying parameters on bacterial inactivation can also be dX E
correlated by comparing drying kinetics and bacterial inactivation kd ¼ k0 1 þ b, exp  d (4)
dt Rg T
using the same drying medium (Fu et al., 2013; Perdana et al., 2013).
Further, the drying rate may be relatively slow when applying mild
    
conditions, thus enabling the analysis of phase transition phe- dT dX E
nomenon such as crystallization, sol-gel transition or skin forma- kd ¼ k0 1 þ a, 1 þ b, exp  d (5)
dt dt Rg T
tion. These observations may offer new avenues for the design of
functional probiotic particles via spray-drying. For example, lactose
crystallization during droplet drying was found to be related to    
dX dX 2 E
bacterial inactivation, which is probably dependent on the crystal kd ¼ k0 1 þ a, þ b, exp  d (6)
dt dt Rg T
shape formed by different media (Perdana et al., 2014). In addition,
the milk protein type and its denaturation degree are important The prediction of these three equations was found to be more
factors in determining powder morphology (Khem, Woo, Small, accurate when compared with the traditional non-rate models Eqs.
Chen, & May, 2015; Rajam, Karthik, Parthasarathi, Joseph, & (2) and (3) (Li, Lin, Chen, Chen, & Pearce, 2006).
Anandharamakrishnan, 2012; Sadek et al., 2013; Soukoulis, As well as these first-order-kinetics-based models, a statistical
Behboudi-Jobbehdar, Yonekura, Parmenter, & Fisk, 2014). This for- model was also developed based on the principle of Weibull dis-
mation of particle morphology was recently found to be a factor tribution (Perdana et al., 2013). This hypothesizes the death of
that triggered bacterial inactivation, probably due to the mechan- microbial cells during spray-drying as probabilities rather than
ical stresses in play (Khem, Bansal, Small, & May, 2016; Wang et al., determinism. There is therefore a distribution of inactivation times,
2016). which means the survival curve of bacteria is a cumulative form of
Based on the single droplet drying technique and the develop- the inactivation times. Temperature and moisture content are
ment of an in situ analytical technique, it is also possible to gain incorporated as two factors determining the inactivation of pro-
insights into the interaction between drying conditions and the biotic Lb. plantarum WCFS1.
structure of bacterial cell components, thus suggesting specific Overall, the predictions of these models agreed well with the
strategies for the protection of bacterial cells, or even cellular experimental data on bacterial inactivation during single droplet
components at the molecular level. drying. However, the predictions cannot be considered precise,
particularly in the context of real-life spray-drying. This may be due
3.5.2. Modelling to the much faster drying rate in spray-drying than in droplet
Based on the “in-process” data obtained using the single droplet drying, and because factors such as oxidative stress, osmotic stress
drying technique, mathematical models have been developed to and the protective capacity of the medium, must also be taken into
predict the inactivation of probiotics during thermal convective account. Furthermore, the intrinsic tolerance of bacteria varies
drying. significantly over different species and strains, which may lead to
The model based on first order kinetics was reviewed by (Chen problems in applying these models to different cases.
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 11

4. Post-drying stage survive better when the aw was lower than 0.33 (Poddar et al.,
2014). It was also found in the context of vacuum-drying that Lb.
4.1. Storage stability of bacteria in powder form paracasei maintained significantly higher survival at an aw of 0.07
when compared with that seen at 0.22 and 0.33 following storage
4.1.1. Injury-induced loss of viability and strain-dependent shelf-life at 20  C and 37  C for two months, respectively (Foerst, Kulozik,
The numbers of viable bacterial cells tend to decrease during Schmitt, Bauer, & Santivarangkna, 2012). Another study per-
storage (generally less than 6 months), and particularly during the formed on Lb. rhamnosus GG also showed that the viability of
early stages (up to one month), possibly due to cell injury during bacteria in dried crushed flaxseed dropped rapidly, with an aw at
spray-drying (Wang et al., 2004). For instance, good survival was 0.43 and 0.22 during long-term storage, while a loss of viability of
found to depend on selecting the best survivors following spray- only 0.29 log10 units was found with an aw at 0.11 (Vesterlund et al.,
drying (Gardiner et al., 2000). The stability of a freeze-dried cul- 2012).
ture has been reported as being better than that of a spray-dried
culture, suggesting the damage caused by heat stress (Wang 4.1.4. Exposure to oxygen: lipid oxidation
et al., 2004). During storage, the oxidation and subsequent saturation of
The storage conditions of powders may also have a significant membrane lipids exert a negative impact on viability (Teixeira,
influence on the survival of probiotics. Key parameters which must Castro, & Kirby, 1996). Changes to the degree of lipid unsatura-
be controlled include: storage temperature, moisture content or tion, which increased over time, were shown to markedly affect the
water activity, exposure to oxygen and light, powder composition passive permeability of the membrane (Teixeira et al., 1996).
and storage materials (Morgan et al., 2006). These conditions may Furthermore, the products of lipid peroxidation have been shown
cause heat, desiccation, oxidative and starvation stress to bacteria, to induce damage to the bacterial cell wall, cell membrane and DNA
which can lead to a considerable loss of the viability of probiotics during storage (Teixeira, Castro, Malcata, 1995).
during long-term storage. The intrinsic tolerance of bacterial strains In order to limit these detrimental effects, the drying medium
is known to play a critical role in overcoming their inactivation due can be supplemented with an antioxidant. However, contradictory
to either spray-drying injury or storage-related adverse stresses. results have been reported in the literature regarding this strategy.
The shelf-life of probiotic powders is therefore dependent on the The addition of ascorbic acid and monosodium glutamate during
bacterial strains concerned. spray-drying improved culture viability during powder storage
(Sunny-Roberts & Knorr, 2009; Teixeira, Castro, & Kirby, 1995),
4.1.2. Storage temperature and glass transition temperature (Tg) although other studies found that they had detrimental effects on
Probiotic viability in a powder is inversely related to storage culture stability during storage. For example, the addition of
temperature. This has been demonstrated extensively in probiotic- ascorbic acid and monosodium glutamate protected Lb. bulgaricus
containing powders obtained from spray-drying (Barbosa et al., cells, but only during storage at 4  C. At 20  C, the death rate of the
2015a; Corcoran et al., 2004; Silva, Carvalho, Teixeira, & Gibbs, culture was even higher in the presence of these compounds than
2002; Wang et al., 2004). However, prolongation of the shelf-life in the control sample (Teixeira, Castro, Malcata, 1995). This could be
of probiotics by reducing the storage temperature is not favored explained by the pro-oxidant properties of ascorbic acid as a metal
by industry because of the cost of chilled or frozen storage, even ions reducer, in addition to its antioxidant function as a radical
though storage at room temperature still poses an overwhelming scavenger.
challenge to the stability of probiotic powders. Barbosa, Borges, and Teixeira (2015b) also reported that the
During storage, the viability of probiotics is affected in particular exposure of Lb. plantarum 299v to sub-lethal thermal, acid or
by temperature in relation to the Tg of the dried sample, which in oxidative stress before spray-drying could enhance bacterial sur-
turn is mainly dependent on its moisture content (Passot, Cenard, vival during storage for 180 days at room temperature. However,
Douania, Tre le
a, & Fonseca, 2012). Indeed, dried biological these sub-lethal stresses did not improve the storage stability of
matrices stored at a temperature below their Tg have been shown to Pediococcus acidilactici HA-6111-2 during the same study, which
display greater stability because of the low molecular mobility and indicates the strain-dependence of this strategy.
reaction kinetics induced by the high viscosity of the glassy state
(Buitink, Van den Dries, Hoekstra, Alberda, & Hemminga, 2000). In 4.1.5. Powder composition and protective agents
this respect, monitoring the moisture content during storage may Spray-drying is a method used at an industrial scale to encap-
be of considerable value to controlling the detrimental influence of sulate various ingredients, including microorganisms. Encapsulated
aging on viability, although this has rarely been reported in the bacterial cells have been widely reported as being more stable than
literature. Another strategy consists in supplementing the growth free cells (De Prisco & Mauriello, 2016). Song, Cho, and Park (2003)
medium with specific components (e.g. high-molecular-weight showed that the protective effect of encapsulation during storage
maltodextrins) in order to increase the overall Tg of the final increased as the storage temperature rose. The effects of encapsu-
dried product and thus better maintain viability during storage lation on the shelf-life of bacteria are also closely linked to the
under given conditions (Conrad et al., 2000). encapsulating carrier (drying medium).
The protective agents used during drying (i.e. efficient drying
4.1.3. Moisture content: the influence of water activity medium) may also protect bacteria during storage. For example,
It is well known that the lower limit aw for bacterial growth is non-reducing disaccharides (trehalose and sucrose) provided good
around 0.6 (Beuchat, 1981). This suggests that a high aw might lead protection for Enterococcus faecium and Lb. plantarum during both
to a reduction of probiotic viability and an increase in the risk of fluidized-bed drying and storage (Strasser, Neureiter, Geppl, Braun,
contamination during storage. The survival rate is not related lin- & Danner, 2009). However, a medium suitable for drying may be
early to aw. In most studies, the optimal range of aw values for the not effective in protecting bacteria during powder storage, and vice
storage of probiotics was lower than 0.2. For example, the relatively versa. For instance, skimmed milk to which polydextrose and oli-
poorer survival of S. thermophilus CCRC 14085 and B. longum B6 gofructose have been added was found to decrease the stability of
detected in spray-dried fermented soymilk after storage could Lb. rhamnosus during powder storage when compared to skimmed
partially be attributed to the high aw of this dried product (Wang milk alone, although this did not affect protection during spray-
et al., 2004). Spray-dried Lb. paracasei CRL 431 was found to drying (Ananta, Volkert, & Knorr, 2005). The addition of chitosan
12 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

to the drying medium was found to decrease viability of Lb. aci- 2004). Furthermore, the reconstitution behavior of freeze-dried
dophilus NCIMB 701748 during spray-drying, but provided excellent yoghurt was better than that of spray-dried yoghurt, which may
protection for the bacterial cells during long-term storage be due to the thermal cellular injury of spray-dried bacteria (Kumar
(Yonekura, Sun, Soukoulis, & Fisk, 2014). & Mishra, 2004).
It is also possible to apply a sub-lethal stress to induce the Regarding the effect of the rehydration medium, Teixeira, Castro,
storage tolerance of probiotics by modifying the drying medium. In & Kirby (1995); Teixeira, Castro, Malcata (1995) demonstrated that
our recent work, a strategy was proposed to utilize hypertonic there were no significant differences between the recovery rates of
sweet whey as both growth and drying medium (Huang et al., dried Lb. bulgaricus when using skimmed milk, MRS broth, deion-
2016a). Its high osmolality triggered the stress tolerance of pro- ized water or phosphate buffer as the rehydration medium. Only
biotics and accumulation of intracellular compatible solutes minor differences (of a maximum 0.5 log CFU mL1) were observed
including trehalose, together with the encapsulation effect pro- between the recovery rates of spray-dried Lb. paracasei NFBC338
vided by thick and adequate solid elements (Fig. 3C). The storage powders in various rehydration media, which included AM buffer
stability of probiotic powders was thus significantly enhanced (0.01 mol/L K2HPO4 and 0.01 mol/KH2PO4), AM buffer containing
during storage. 20% (w/v) sucrose, maximum recovery diluent (MRD), 10% (w/v)
RSM, and sterile water (Desmond et al., 2001).
4.1.6. Storage materials
Options for packaging include different types of barriers to the 4.2.2. Changes to bacterial metabolism
aforementioned reactive agents such as humidity, oxygen and light. Unlike freeze-drying, spray-drying has been found to delay
High barrier plastic bags and blister packs have been reported lactic acid production in Lc. cremoris, Lactobacillus casei ssp. pseu-
(Morgan et al., 2006). It was found that laminated pouches ensured doplantarum and S. thermophilus (To & Etzel, 1997). The higher the
better protection during the storage of S. thermophilus CCRC 14085 outlet air temperatures, the longer was the lag time before the
and B. longum B6, when compared to glass or PET bottles. production of acid. For example, the lag time for Lb. pseudoplanta-
Modifying the composition of the surrounding atmosphere by rum was shorter (2 h) at Toutlet 65  C than at Toutlet 90  C (10 h).
means of a vacuum or nitrogen generally enhances the storage Apart from its influence on lactic acid fermentation, more attention
stability of foods. However, the improvement to the storage sta- needs to be paid to how lag time affects the beneficial effects of
bility of probiotic powders procured using this method appears to spray-dried probiotics. For instance, it has been documented that
be slight or insignificant (Clementi & Rossi, 1984; Espina & Packard, bacteria retain their ability to produce bacteriocin peptides
1979; Huang et al., 2017). Ambiguous results have thus been ob- following spray-drying: this has been shown for nisin (Nisaplin®;
tained regarding the effects of the storage atmosphere on bacterial Danisco A/S, Copenhagen, Denmark), lacticin 3147 (Morgan et al.,
stability; so they are rarely reported. 2006), and other bacteriocins produced by both lactobacilli and
lactococci (Mauriello, Aponte, Andolfi, Moschetti, & Villani, 1999;
4.2. Applications of dried bacteria Zhang, Luan, Zhang, Zhang, & Hao, 2015). Pe rez-Chabela, Lara-
Labastida, Rodriguez-Huezo, and Totosaus (2013) reported that
4.2.1. Rehydration capacity the incorporation of spray-dried LAB in meat batters enhanced the
After drying, a number of functionalities are expected to be initial LAB population and a concomitant reduction of Enter-
retained in probiotics in powders, such as metabolic activity, obacteria levels in the meat was seen during storage. Lb. salivarius
tolerance towards human gastrointestinal juices, adherence to UCC 118 was also reported to retain its ability to produce bacte-
epithelial surfaces, antagonistic activity against pathogens and riocin after spray-drying, even at high outlet air temperatures (up
immunoregulatory capacities. To operate these probiotic func- to 95  C) (Gardiner et al., 2000). In another study, the spray-drying
tionalities, rehydration is generally the first step for products in a process did not affect the antagonistic activity of Lb. sakei and Lb.
powder form. salivarius against Staphylococcus aureus, Listeria innocua and Listeria
The recovery of spray-dried probiotics improved after the monocytogenes (Silva et al., 2002).
rehydration temperature was increased, and was seen to be a Overall, the residual viability of probiotics after spray-drying is
strain-dependent process (Mille, Obert, Beney, & Gervais, 2004). At an important indicator when characterizing probiotic activity
the same time, this temperature should not be as high as the heat- within spray dried powders. Low viability is more a reflection of the
lethal temperature (i.e. 50  C for most LAB). Rehydration of spray cellular injury of probiotics during the drying process. Severe
dried S. thermophilus and B. longum resulted in enhanced recovery cellular injury may significantly influence the lag time of bacterial
of live cells as much as the temperature increased, within the range growth and primary metabolism-related functionalities, such as
5e50  C (Wang et al., 2004). When the rehydration of Lb. bulgaricus lactic acid production. However, when regrowth reaches the late
was performed at 4  C, the cell concentration was only log phase or early stationary phase, the culture contains a sufficient
2.1  109 CFU g1, whereas it reached 1.5  1010 CFU g1 at 30  C or bacterial population and new generations of the bacteria are not
37  C. Between these values, the live bacteria recovery increased influenced by the drying process, thus leading to them being less
steadily in line with the rehydration temperature: 3  109 CFU g1, influenced by the secondary metabolism-related functionalities of
3.5  109 CFU g1, 5.2  109 CFU g1, and 1.1  1010 CFU g1 at 10  C, bacteria such as bacteriocin-producing ability and enzyme activity.
15  C, 20  C and 25  C, respectively (Mille et al., 2004). Conversely,
temperature had no significant effect on the final bacterial con- 4.2.3. Resistance to food manufacturing processes
centration when dried Lb. plantarum samples were rehydrated at Due to the convenience of powder-form products, spray-dried
30  C or 37  C (1.5  1010 CFU g1 and 1.3  1010 CFU g1, respec- bacteria can be expected to be used as an ingredient to be added
tively). This could be explained by the higher cellular resistance of in other foods. This often necessitates the heat resistance of spray-
Lb. plantarum. dried probiotics because of the presence of a thermal process in
Interestingly, the difference in the optimum rehydration tem- most food manufacturing (Dianawati, Mishra, & Shah, 2015; Zhang,
perature observed when recovering freeze-dried (20  C) and spray- Huang, Ananingsih, Zhou, & Chen, 2014). Spray-dried bacteria
dried (35  Ce50  C) bacteria further demonstrated that a physio- generally display higher heat resistance when compared to free
logical difference does exist between freeze-dried and spray-dried bacteria. For example, the survival of spray-dried Lactobacillus
cells of S. thermophilus CCRC 14085 and B. longum B6 (Wang et al., reuteri DSM 17938 (with co-cross-linked alginate and chitosan as
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 13

drying medium) was significantly better than that of free bacteria 2015). Therefore, the results of simple in vitro assays cannot be
during the baking of a chocolate souffle  (Malmo, Storia, & extrapolated directly to living organisms. It is problematic to
Mauriello, 2013). The reduction in the bacterial population of perform numerous digestion trials in the context of ethically and
spray-dried Lb. plantarum BM-1 was only 0.08 log CFU mL1 after technically challenging animal and human studies. The novel dy-
exposure to 60  C, whereas the reduction in free cells reached 2.58 namic gastrointestinal models now available, which can better
log CFU mL1 (Zhang et al., 2015). This can be explained by three reflect the complex digestive systems of living organisms, now
mechanisms: first, the encapsulation effect of the drying medium need to be applied to testing the behavior during digestion and
(De Prisco & Mauriello, 2016; Sabikhi, Babu, Thompkinson, & efficiency of delivery of probiotic products (Chen et al., 2016;
Kapila, 2010); second, the lower water activity surrounding bacte- Cordonnier et al., 2015).
rial cells (Huang et al., 2017; Wang et al., 2016), and third, the
possible induction of a heat shock response due to the heat stress 4.2.5. Adhesion to epithelial cells
experienced during spray-drying (Rossi, Zotta, Iacumin, & Reale, The importance of the adhesion of probiotics to epithelial cells
2016). In addition to the thermal process, probiotics or LAB cells or the intestinal mucus has been well documented. However, very
(which are used as a starter) are often exposed to an acidic envi- few publications have reported the effects of spray-drying on the
ronment in foods because of the drop in pH during fermentation. adhesion capacity of probiotics to intestinal epithelial cells. Lb.
Spray-dried bacteria can often survive better in an acidic environ- plantarum 83114 and Lactobacillus kefir 8321 did not lose their
ment during their shelf-life. For example, Dimitrellou et al. (2016) ability to adhere to intestinal Caco-2/TC-7 cells, while Lb. kefir 8348
showed that spray-dried Lb. casei ATCC393 had an increased sur- displayed a significant loss of adhesion capacity after spray-drying
vival rate in fermented milk during refrigerated storage, by com- under the same conditions (Golowczyc, Silva, Teixeira, De Antoni, &
parison with free bacteria. Abraham, 2011). This result is in agreement with the sensitivity of
However, it is also possible that the cellular injury caused during bacterial strains to spray-drying. Lb. kefir 8348 had the poorest
spray-drying may lead to a higher bacterial death rate during survival after spray-drying (~30%), while the survival of Lb. plan-
subsequent processing (Wu, 2008). Hence, the adequate bacterial tarum 83114 and Lb. kefir 8321 was between 80% and 90%.
viability and activity in spray-dried powders constitute important Improved survival (approximately higher than 50%) may guar-
criteria for food applications. antee the retention of the adhesion capacity of probiotics because
of the insignificant injury of cell surface structures. Further, pro-
4.2.4. Resistance to gastrointestinal conditions biotics adhesion capacity may also be improved after spray-drying
Survival is essential for bacteria that are targeted to populate the by using prebiotics in the drying medium (Brink, Todorov, Martin,
human gut: this is one of the most important issues when pro- Senekal, & Dicks, 2006).
curing health benefits with probiotics (Anselmo, McHugh, Webster,
Langer, & Jaklenec, 2016; De Prisco & Mauriello, 2016). 4.2.6. Immunomodulation ability
The challenge is to maintain probiotic viability during delivery Immunomodulation is one of the most important probiotic
through the digestive tract. More specifically, a considerable loss of functionalities and has been extensively investigated during the
viability was observed when bacterial cells were exposed to acid past decade. However, there have been very few studies which
and bile stresses, gastric or intestinal enzymes, and a mechanical reported the effects of spray-drying on the efficacy of probiotic
shearing force during consumption. immunomodulation. The effects of air drying, freeze-drying and
In vitro studies have shown that an appropriate spray-drying spray-drying on the immunomodulation ability of the probiotics
medium can protect probiotics against stress during digestion. Lb. plantarum CNRZ 1997 and Lactobacillus zeae CNRZ 2268 and
For example, a loss of viability was limited by using gum acacia (GA) B. bifidum CIP 56.7 were compared with peripheral blood mono-
as a carrier during spray-drying: the viability of GA-treated bacteria nuclear cells (PBMC) in vitro. The probiotic functionality was not
was 100-fold higher than the control following exposure for directly linked to cell survival, and displayed strain-dependent
120 min to porcine gastric juice at 37  C (Desmond et al., 2002). sensitivity to each of the drying methods. After different drying
Such protection may be reliant on the resistance of GA to digestion processes, the probiotic powders might positively or negatively
(Arslan, Erbas, Tontul, & Topuz, 2015). As well as GA, it has also been modify the bacterial immunomodulation capacity, and spray-
suggested that a dairy matrix might function as a buffering agent, drying appeared to be the best drying process because of its ef-
thereby protecting ingested bacteria during transit through the fect on decreasing the production of PBMC IL-12 (Iaconelli et al.,
upper gastrointestinal tract (Würth et al., 2015). The anionic poly- 2015). Furthermore, during an in vivo study performed with Lb.
saccharide alginate, and the cationic polysaccharide chitosan, are acidophilus A9, Lb. paracasei A13 and Lb. casei Nad, spray-drying
also known for their biocompatible resistance to digestion and was carried out using 20% (w/v) skimmed milk as the drying
their mucoadhesive properties (Anselmo et al., 2016; Kim et al., medium, and the probiotic powders were administered to mice for
2014). Rajam et al. (2012) showed that a combination of sodium 5 and 10 days. A significantly higher number of Immunoglobulin A
alginate and denatured WPI as the drying medium could improve (IgA)-producing cells in the small intestine were induced by spray-
the survival rate and controlled core release behavior of spray-dried dried cultures when compared with fresh cultures (Pa ez et al.,
Lb. plantarum during simulated acidic and bile conditions. By 2013).
comparison with sodium alginate, chitosan was reported to achieve Immunomodulation by probiotics is mainly due to molecular
better protection of Bifidobacterium breve during fluid-bed drying, interactions between the surfaces of the bacterial cell and host cell
and of Lb. acidophilus during simulated digestion (Cook, Tzortzis, (Bron, Van Baarlen, & Kleerebezem, 2011). The spray-dried pro-
Charalampopoulos, & Khutoryanskiy, 2011; Yonekura et al., 2014). biotics are encapsulated by the drying medium. The barrier effect of
This could be explained by the better electrostatic interactions the drying medium may therefore modify the interaction between
between chitosan (cationic polysaccharide) and bacterial cells the microbes and the host, and this effect is probably dependent on
(negative surface charge) (Anselmo et al., 2016). the digestion behavior of the matrix. The advantage of spray-drying
To date, only a few studies have been performed in vivo. It has probiotics may be linked to the barrier effect of the drying medium
been shown that milk-protein-based microcapsules display a pro- in that it can protect the bacterial surface structure from the
tective effect on probiotics during simulated in vitro digestion, but digestion process, including pH-induced conformation changes,
not in the mouse gastrointestinal system (in vivo) (Würth et al., enzyme catalysis and the shearing force of gastrointestinal tracts.
14 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

However, severe cell damage caused by spray-drying may also lead iii. Most of the studies cited here were performed on laboratory-
to a diminished adhesion capacity of bacteria, heat or osmotic- scale dryers. Scaling-up to industrial conditions, when
induced conformation changes to the surface structure or a shield tested, was not always successful. This key issue should be
effect of the drying medium. However, this remains a hypothesis to addressed in future trials to cover all aspects from molecular
be addressed by future research. mechanisms to industrial production.
iv. The stability of probiotic powders during storage was char-
5. Conclusion and perspectives acterized by an interval test of viability in most studies, while
changes to functionality were rarely reported. A clearer un-
The food and pharmaceutical industries are currently faced with derstanding of the physiological mechanism underlying a
new opportunities and new challenges with respect to the drying of loss of viability, and its relationship with storage conditions,
probiotics by spray-drying. Some key messages can be drawn from would constitute a major breakthrough. The functionality of
the present study: probiotic powders needs to be validated alongside their
viability. This requires more experimental data from both
i. The growth and drying processes should be adapted and in vitro and in vivo studies. In terms of in vitro studies, more
optimized for each probiotic strain; “near real” dynamic and simulated digestive systems should
ii. Triggering bacterial stress adaptation before drying is a be used.
strategy that can improve probiotic viability during drying v. Future attempts should target the more challenging and
and storage, where heat, oxidative and osmotic adaptations fragile “next generation” probiotics, such as Akkermansia
are closely related to survival following spray-drying; muciniphila.
iii. Harvesting cells in the stationary phase, and neutralizing the
pH before drying, are generally beneficial; To conclude, a clearer understanding of bacterial physiology
iv. A protective drying medium can increase the resistance of during growth, drying and storage, including strain-specificity and
probiotics to spray-drying, storage and gastrointestinal industrial conditions, will be determinant in achieving advances in
conditions; the development of stable and efficient probiotic products.
v. A reduction in the outlet air temperature can be achieved by
process optimization in order to reduce bacterial death, Acknowledgements
although powder quality should also be taken into account;
vi. A single droplet drying technique is an effective tool to The authors offer their sincere thanks to the joint PhD project
monitor and optimize the viability of probiotics during initiated between Agrocampus Ouest and Soochow University.
spray-drying;
vii. Maintaining probiotic viability during storage can be ach- References
ieved by lowering the storage temperature, controlling the
powder moisture content and avoiding exposure to oxygen; Abe, F., Miyauchi, H., Uchijima, A., Yaeshima, T., & Iwatsuki, K. (2009). Effects of
viii. Cellular injury caused by spray-drying leads to a decreased storage temperature and water activity on the survival of bifidobacteria in
powder form. International Journal of Dairy Technology, 62(2), 234e239.
stability of probiotics during storage, an extended lag time Alcantara, C., Revilla-Guarinos, A., & Zuniga, M. (2011). Influence of two-component
and diminished lactic acid production during regrowth; signal transduction systems of Lactobacillus casei BL23 on tolerance to stress
ix. Residual probiotic viability is an important indicator to conditions. Applied and Environmental Microbiology, 77(4), 1516e1519. https://
doi.org/10.1128/AEM.02176-10.
characterize the functionality of probiotic powders. Survival Ananta, E., & Knorr, D. (2003). Pressure-induced thermotolerance of Lactobacillus
above 50% seems to be a safe range to guarantee this rhamnosus GG. Food Research International, 36(9e10), 991e997. https://doi.org/
functionality. 10.1016/j.foodres.2003.07.001.
Ananta, E., Volkert, M., & Knorr, D. (2005). Cellular injuries and storage stability of
spray-dried Lactobacillus rhamnosus GG. International Dairy Journal, 15(4),
However, we also identified some deficiencies and contradictory 399e409. http://doi.org/10.1016/j.idairyj.2004.08.004.
results in the literature. Further knowledge and research are Anselmo, A. C., McHugh, K. J., Webster, J., Langer, R., & Jaklenec, A. (2016). Layer-by-
layer encapsulation of probiotics for delivery to the microbiome. Advanced
required to address the following topics:
Materials, 28, 9486e9490. https://doi.org/10.1002/adma.201603270.
Arslan, S., Erbas, M., Tontul, I., & Topuz, A. (2015). Microencapsulation of probiotic
i. Single droplet drying can offer insights into the droplet- Saccharomyces cerevisiae var. boulardii with different wall materials by spray
particle conversion process under defined conditions and drying. LWT - Food Science and Technology, 63(1), 685e690. https://doi.org/10.
1016/j.lwt.2015.03.034.
consuming much less material, energy and time. One major Baker-Austin, C., & Dopson, M. (2007). Life in acid: pH homeostasis in acidophiles.
challenge is identification of the molecular mechanisms Trends in Microbiology, 15(4), 165e171. https://doi.org/10.1016/j.tim.2007.02.
responsible for bacterial injury or adaptation. The maximal 005.
Barbosa, J., Borges, S., Amorim, M., Pereira, M. J., Oliveira, A., Pintado, M. E., et al.
residual viability of probiotics could be achieved by either (2015a). Comparison of spray drying, freeze drying and convective hot air
mapping the optimal drying parameters or designing pro- drying for the production of a probiotic orange powder. Journal of Functional
tective media. Foods, 17, 340e351. https://doi.org/10.1016/j.jff.2015.06.001.
Barbosa, J., Borges, S., & Teixeira, P. (2015b). Influence of sub-lethal stresses on the
ii. Greater attention should be paid to the links between survival of lactic acid bacteria after spray-drying in orange juice. Food Micro-
different units, and in particular the links between probiotic biology, 52, 77e83. https://doi.org/10.1016/j.fm.2015.06.010.
growth and drying, and between drying and powder Barbosa, J., & Teixeira, P. (2016). Development of probiotic fruit juice powders by
spray drying: A review. Food Reviews International, 33(4), 335e358. https://doi.
collection. More specifically, and unlike freeze-drying, pro- org/10.1080/87559129.2016.1175016.
biotics may remain for hours in the liquid feed concentrate Beuchat, L. R. (1981). Microbial stability as affected by water activity. Cereal Foods
before being pumped to the dryer. During this period, the World, 26(7), 345e349.
Bielecka, M., & Majkowska, A. (2000). Effect of spray drying temperature of yoghurt
emergence of quorum sensing, stress adaptation, or the
on the survival of starter cultures, moisture content and sensoric properties of
growth or death of probiotics may be possible. In addition, yoghurt powder. Nahrung-Food, 44(4), 257e260.
the link between spray-drying and powder collection is not Bimbenet, J. J., Schuck, P., Roignant, M., Brule , G., & Me
jean, S. (2002). Heat balance
often addressed by many studies. For example, the time of a multistage spray-dryer: Principles and example of application. Lait, 82(4),
541e551. https://doi.org/10.1051/lait:2002031.
required and method used to remove powders attached to Bove, P., Capozzi, V., Garofalo, C., Rieu, A., Spano, G., & Fiocco, D. (2012). Inactivation
the dryer wall, and to collect them, need further study. of the ftsH gene of Lactobacillus plantarum WCFS1: Effects on growth, stress
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 15

tolerance, cell surface properties and biofilm formation. Microbiological Desmond, C., Stanton, C., Fitzgerald, G. F., Collins, K., & Ross, R. P. (2001). Environ-
Research, 167(4), 187e193. https://doi.org/10.1016/j.micres.2011.07.001. mental adaptation of probiotic lactobacilli towards improvement of perfor-
Brink, M., Todorov, S. d., Martin, J. h., Senekal, M., & Dicks, L. m. t. (2006). The effect mance during spray drying. International Dairy Journal, 11(10), 801e808.
of prebiotics on production of antimicrobial compounds, resistance to growth at Dianawati, D., Mishra, V., & Shah, N. P. (2015). Survival of microencapsulated pro-
low pH and in the presence of bile, and adhesion of probiotic cells to intestinal biotic bacteria after processing and during storage: A review. Critical Reviews in
mucus. Journal of Applied Microbiology, 100(4), 813e820. https://doi.org/10.1111/ Food Science and Nutrition, 56(10), 1685e1716. https://doi.org/10.1080/
j.1365-2672.2006.02859.x. 10408398.2013.798779.
Broeckx, G., Vandenheuvel, D., Claes, I. J. J., Lebeer, S., & Kiekens, F. (2016). Drying Dijkstra, A. R., Setyawati, M. C., Bayjanov, J. R., Alkema, W., Van Hijum, S. A. F. T.,
techniques of probiotic bacteria as an important step towards the development Bron, P. A., et al. (2014). Diversity in robustness of Lactococcus lactis strains
of novel pharmabiotics. International Journal of Pharmaceutics, 505(1e2), during heat stress, oxidative stress, and spray drying stress. Applied and Envi-
303e318. https://doi.org/10.1016/j.ijpharm.2016.04.002. ronmental Microbiology, 80(2), 603e611. https://doi.org/10.1128/AEM.03434-13.
Bron, P. A., Van Baarlen, P., & Kleerebezem, M. (2011). Emerging molecular insights Dimitrellou, D., Kandylis, P., Petrovi c, T., Dimitrijevic-Brankovi c, S., Levic, S.,
into the interaction between probiotics and the host intestinal mucosa. Nature Nedovi c, V., et al. (2016). Survival of spray dried microencapsulated Lactoba-
Reviews Microbiology. https://doi.org/10.1038/nrmicro2690. cillus casei ATCC 393 in simulated gastrointestinal conditions and fermented
Buitink, J., Van den Dries, I. J., Hoekstra, F. A., Alberda, M., & Hemminga, M. A. milk. LWT - Food Science and Technology, 71, 169e174. https://doi.org/10.1016/j.
(2000). High critical temperature above Tg may contribute to the stability of lwt.2016.03.007.
biological systems. Biophysical Journal, 79(2), 1119e1128. Eratte, D., McKnight, S., Gengenbach, T. R., Dowling, K., Barrow, C. J., & Adhikari, B. P.
Bustamante, M., Oomah, B. D., Rubilar, M., & Shene, C. (2017). Effective Lactobacillus (2015). Co-encapsulation and characterisation of omega-3 fatty acids and pro-
plantarum and Bifidobacterium infantis encapsulation with chia seed (Salvia biotic bacteria in whey protein isolateegum Arabic complex coacervates. Jour-
hispanica L.) and flaxseed (Linum usitatissimum L.) mucilage and soluble protein nal of Functional Foods, 19(Part B), 882e892. https://doi.org/10.1016/j.jff.2015.01.
by spray drying. Food Chemistry, 216, 97e105. https://doi.org/10.1016/j. 037.
foodchem.2016.08.019. Espina, F., & Packard, V. S. (1979). Survival of Lactobacillus acidophilus in a spray-
Bustamante, M., Villarroel, M., Rubilar, M., & Shene, C. (2015). Lactobacillus aci- drying process. Journal of Food Protection, 42(2), 149e152.
dophilus La-05 encapsulated by spray drying: Effect of mucilage and protein Ferrando, V., Quiberoni, A., Reinhemer, J., & Su arez, V. (2015). Resistance of func-
from flaxseed (Linum usitatissimum L.). LWT - Food Science and Technology, 62(2), tional Lactobacillus plantarum strains against food stress conditions. Food
1162e1168. https://doi.org/10.1016/j.lwt.2015.02.017. Microbiology, 48, 63e71. https://doi.org/10.1016/j.fm.2014.12.005.
Chen, X. D., & Patel, K. C. (2007). Micro-organism inactivation during drying of small Ferreira, V., Soares, V., Santos, C., Silva, J., Gibbs, P. A., & Teixeira, P. (2005). Survival
droplets or thin-layer slabs - a critical review of existing kinetics models and an of Lactobacillus sakei during heating, drying and storage in the dried state when
appraisal of the drying rate dependent model. Journal of Food Engineering, 82(1), growth has occurred in the presence of sucrose or monosodium glutamate.
1e10. https://doi.org/10.1016/j.jfoodeng.2006.12.013. Biotechnology Letters, 27(4), 249e252. https://doi.org/10.1007/s10529-004-
Chen, L., Xu, Y., Fan, T., Liao, Z., Wu, P., Wu, X., et al. (2016). Gastric emptying and 8351-x.
morphology of a “near real” in vitro human stomach model (RD-IV-HSM). Foerst, P., & Kulozik, U. (2011). Modelling the dynamic inactivation of the probiotic
Journal of Food Engineering, 183, 1e8. https://doi.org/10.1016/j.jfoodeng.2016.02. bacterium L. paracasei ssp. paracasei during a low-temperature drying process
025. based on stationary data in concentrated systems. Food and Bioprocess Tech-
Clementi, F., & Rossi, J. (1984). Effect of drying and storage-conditions on survival of nology, 5(6), 2419e2427. https://doi.org/10.1007/s11947-011-0560-4.
leuconostoc oenos. American Journal of Enology and Viticulture, 35(3), 183e186. Foerst, P., Kulozik, U., Schmitt, M., Bauer, S., & Santivarangkna, C. (2012). Storage
Coldewey, S. M., Hartmann, M., Schmidt, D. S., Engelking, U., Ukena, S. N., & stability of vacuum-dried probiotic bacterium Lactobacillus paracasei F19. Food
Gunzer, F. (2007). Impact of the rpoS genotype for acid resistance patterns of and Bioproducts Processing, 90(2), 295e300. https://doi.org/10.1016/j.fbp.2011.
pathogenic and probiotic Escherichia coli. BMC Microbiology, 7(1), 21. https://doi. 06.004.
org/10.1186/1471-2180-7-21. Fu, N., & Chen, X. D. (2011). Towards a maximal cell survival in convective thermal
Conrad, P. B., Miller, D. P., Cielenski, P. R., & de Pablo, J. J. (2000). Stabilization and drying processes. Food Research International, 44(5), 1127e1149. https://doi.org/
preservation of Lactobacillus acidophilus in saccharide matrices. Cryobiology, 10.1016/j.foodres.2011.03.053.
41(1), 17e24. https://doi.org/10.1006/cryo.2000.2260. Fu, W. Y., Suen, S. Y., & Etzel, M. R. (1995). Inactivation of Lactococcus lactis ssp. lactis
Cook, M. T., Tzortzis, G., Charalampopoulos, D., & Khutoryanskiy, V. V. (2011). Pro- C2 and alkaline phosphatase during spray drying. Drying Technology, 13(5e7),
duction and evaluation of dry alginate-chitosan microcapsules as an enteric 1463e1476. https://doi.org/10.1080/07373939508917033.
delivery vehicle for probiotic bacteria. Biomacromolecules, 12(7), 2834e2840. Fu, N., Woo, M. W., Selomulya, C., & Chen, X. D. (2013). Inactivation of Lactococcus
https://doi.org/10.1021/bm200576h. lactis ssp. cremoris cells in a droplet during convective drying. Biochemical En-
Corcoran, B. M., Ross, R. P., Fitzgerald, G. F., Dockery, P., & Stanton, C. (2006). gineering Journal, 79, 46e56. https://doi.org/10.1016/j.bej.2013.06.015.
Enhanced survival of GroESL-overproducing Lactobacillus paracasei NFBC 338 Gardiner, G. E., O'Sullivan, E., Kelly, J., Auty, M. A., Fitzgerald, G. F.,
under stressful conditions induced by drying. Applied and Environmental Collins, J. K., … Stanton, C. (2000). Comparative survival rates of human-derived
Microbiology, 72(7), 5104e5107. https://doi.org/10.1128/AEM.02626-05. probiotic Lactobacillus paracasei and L. salivarius strains during heat treatment
Corcoran, B. M., Ross, R. P., Fitzgerald, G. F., & Stanton, C. (2004). Comparative and spray drying. Applied and Environmental Microbiology, 66(6), 2605e2612.
survival of probiotic lactobacilli spray-dried in the presence of prebiotic sub- Ghandi, A., Powell, I. B., Howes, T., Chen, X. D., & Adhikari, B. (2012). Effect of shear
stances. Journal of Applied Microbiology, 96(5), 1024e1039. rate and oxygen stresses on the survival of Lactococcus lactis during the at-
Cordonnier, C., The venot, J., Etienne-Mesmin, L., Denis, S., Alric, M., Livrelli, V., et al. omization and drying stages of spray drying: A laboratory and pilot scale study.
(2015). Dynamic in vitro models of the human gastrointestinal tract as relevant Journal of Food Engineering, 113(2), 194e200. https://doi.org/10.1016/j.jfoodeng.
tools to assess the survival of probiotic strains and their interactions with gut 2012.06.005.
microbiota. Microorganisms, 3(4), 725e745. https://doi.org/10.3390/ Golowczyc, M. A., Silva, J., Teixeira, P., De Antoni, G. L., & Abraham, A. G. (2011).
microorganisms3040725. Cellular injuries of spray-dried Lactobacillus spp. isolated from kefir and their

Corona-Hernandez, R. I., Alvarez-Parrilla, E., Lizardi-Mendoza, J., Islas-Rubio, A. R., impact on probiotic properties. International Journal of Food Microbiology,
de la Rosa, L. A., & Wall-Medrano, A. (2013). Structural stability and viability of 144(3), 556e560. https://doi.org/10.1016/j.ijfoodmicro.2010.11.005.
microencapsulated probiotic bacteria: A review: Encapsulation of probiotic Guchte, M., Van de, Serror, P., Chervaux, C., Smokvina, T., Ehrlich, S. D., & Maguin, E.
bacteria. Comprehensive Reviews in Food Science and Food Safety, 12(6), 614e628. (2002). Stress responses in lactic acid bacteria. Antonie van Leeuwenhoek,
http://doi.org/10.1111/1541-4337.12030. 82(1e4), 187e216. https://doi.org/10.1023/A:1020631532202.
Cousin, F. J., Mater, D. D. G., Foligne, B., & Jan, G. (2010). Dairy propionibacteria as Guerin, J., Petit, J., Burgain, J., Borges, F., Bhandari, B., Perroud, C., … Gaiani, C. (2017).
human probiotics: A review of recent evidence. Dairy Science & Technology, Lactobacillus rhamnosus GG encapsulation by spray-drying: Milk proteins clot-
91(1), 1e26. https://doi.org/10.1051/dst/2010032. ting control to produce innovative matrices. Journal of Food Engineering, 193,
Crowe, J. H., Carpenter, J. P., & Crowe, L. M. (1998). The role of vitrification in 10e19. https://doi.org/10.1016/j.jfoodeng.2016.08.008.
anhydrobiosis. Annual Review of Physiology, 60, 73e103. Hill, C., Guarner, F., Reid, G., Gibson, G. R., Merenstein, D. J., Pot, B., … Sanders, M. E.
Crowe, J. H., Crowe, L. M., & Chapman, D. (1984). Preservation of membranes in (2014). Expert consensus document: The International Scientific Association for
anhydrobiotic organisms: The role of trehalose. Science, 223(4637), 701e703. Probiotics and Prebiotics consensus statement on the scope and appropriate
https://doi.org/10.1126/science.223.4637.701. use of the term probiotic. Nature Reviews Gastroenterology & Hepatology, 11(8),
De Angelis, M., Calasso, M., Cavallo, N., Di Cagno, R., & Gobbetti, M. (2016). Func- 506e514. https://doi.org/10.1038/nrgastro.2014.66.
tional proteomics within the genus Lactobacillus. Proteomics, 16(6), 946e962. Huang, S., Cauty, C., Dolivet, A., Le Loir, Y., Chen, X. D., Schuck, P., … Jeantet, R.
https://doi.org/10.1002/pmic.201500117. (2016a). Double use of highly concentrated sweet whey to improve the biomass
De Prisco, A., & Mauriello, G. (2016). Probiotication of foods: A focus on microen- production and viability of spray-dried probiotic bacteria. Journal of Functional
capsulation tool. Trends in Food Science & Technology, 48, 27e39. https://doi.org/ Foods, 23, 453e463. https://doi.org/10.1016/j.jff.2016.02.050.
10.1016/j.tifs.2015.11.009. Huang, S., Me jean, S., Rabah, H., Dolivet, A., Le Loir, Y., Chen, X. D., … Schuck, P.
Desmond, C., Fitzgerald, G. F., Stanton, C., & Ross, R. P. (2004). Improved stress (2017). Double use of concentrated sweet whey for growth and spray drying of
tolerance of GroESL-overproducing Lactococcus lactis and probiotic Lactobacillus probiotics: Towards maximal viability in pilot scale spray dryer. Journal of Food
paracasei NFBC 338. Applied and Environmental Microbiology, 70(10), Engineering, 196, 11e17. https://doi.org/10.1016/j.jfoodeng.2016.10.017.
5929e5936. https://doi.org/10.1128/AEM.70.10.5929-5936.2004. Huang, S., Rabah, H., Jardin, J., Briard-Bion, V., Parayre, S., Maillard, M.-B., … Jan, G.
Desmond, C., Ross, R. P., O'Callaghan, E., Fitzgerald, G., & Stanton, C. (2002). (2016b). Hyperconcentrated sweet whey: A new culture medium that enhances
Improved survival of Lactobacillus paracasei NFBC 338 in spray-dried powders Propionibacterium freudenreichii stress tolerance. Applied and Environmental
containing gum acacia. Journal of Applied Microbiology, 93(6), 1003e1011. Microbiology, 82(15), 4641e4651. https://doi.org/10.1128/AEM.00748-16.
16 S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17

Huang, S., Yang, Y., Fu, N., Qin, Q., Zhang, L., & Chen, X. D. (2014). Calcium-aggre- 302e310. https://doi.org/10.1016/j.ijfoodmicro.2016.09.023.
gated milk: A potential new option for improving the viability of lactic acid Maciel, G. M., Chaves, K. S., Grosso, C. R. F., & Gigante, M. L. (2014). Microencap-
bacteria under heat stress. Food and Bioprocess Technology, 7(11), 3147e3155. sulation of Lactobacillus acidophilus La-5 by spray-drying using sweet whey and
https://doi.org/10.1007/s11947-014-1331-9. skim milk as encapsulating materials. Journal of Dairy Science, 97(4), 1991e1998.
Hussain, M. a., Knight, M. i., & Britz, M. l. (2009). Proteomic analysis of lactose- https://doi.org/10.3168/jds.2013-7463.
starved Lactobacillus casei during stationary growth phase. Journal of Malmo, C., Storia, A. L., & Mauriello, G. (2013). Microencapsulation of Lactobacillus
Applied Microbiology, 106(3), 764e773. https://doi.org/10.1111/j.1365-2672. reuteri DSM 17938 cells coated in alginate beads with chitosan by spray drying
2008.03961.x. to use as a probiotic cell in a chocolate Souffle . Food and Bioprocess Technology,
Iaconelli, C., Lemetais, G., Kechaou, N., Chain, F., Bermúdez-Humara n, L. G., 6(3), 795e805. https://doi.org/10.1007/s11947-011-0755-8.
Langella, P., … Beney, L. (2015). Drying process strongly affects probiotics Mauriello, G., Aponte, M., Andolfi, R., Moschetti, G., & Villani, F. (1999). Spray-drying
viability and functionalities. Journal of Biotechnology, 214, 17e26. https://doi. of bacteriocin-producing lactic acid bacteria. Journal of Food Protection, 62(7),
org/10.1016/j.jbiotec.2015.08.022. 773e777.
Jewell, J. B., & Kashket, E. R. (1991). Osmotically regulated transport of proline by Meerdink, G., & Riet, K. Van ’t (1995). Prediction of product quality during spray
Lactobacillus acidophilus IFO 3532. Applied and Environmental Microbiology, drying. Food and Bioproducts Processing, 73(C), 165e170.
57(10), 2829e2833. Mille, Y., Beney, L., & Gervais, P. (2005). Compared tolerance to osmotic stress in
Kearney, N., Meng, X. C., Stanton, C., Kelly, J., Fitzgerald, G. F., & Ross, R. P. (2009). various microorganisms: Towards a survival prediction test. Biotechnology and
Development of a spray dried probiotic yoghurt containing Lactobacillus para- Bioengineering, 92(4), 479e484.
casei NFBC 338. International Dairy Journal, 19(11), 684e689. https://doi.org/10. Mille, Y., Obert, J. P., Beney, L., & Gervais, P. (2004). New drying process for lactic
1016/j.idairyj.2009.05.003. bacteria based on their dehydration behavior in liquid medium. Biotechnology
Kets, E. P. W., & Bont, J. A. M. D. (1997). Effect of carnitines on Lactobacillus plan- and Bioengineering, 88(1), 71e76.
tarum subjected to osmotic stress. FEMS Microbiology Letters, 146(2), 205e209. Morgan, C. A., Herman, N., White, P. A., & Vesey, G. (2006). Preservation of micro-
https://doi.org/10.1111/j.1574-6968.1997.tb10194.x. organisms by drying: A review. Journal of Microbiological Methods, 66(2),
Kets, E. P. W., Teunissen, P. J. M., & Debont, J. A. M. (1996). Effect of compatible 183e193. https://doi.org/10.1016/j.mimet.2006.02.017.
solutes on survival of lactic acid bacteria subjected to drying. Applied and Nierhaus, K. H. (2014). Mg2þ, Kþ, and the ribosome. Journal of bacteriology, 196(22),
Environmental Microbiology, 62(1), 259e261. 3817e3819. https://doi.org/10.1128/JB.02297-14.
Khem, S., Bansal, V., Small, D. M., & May, B. K. (2016). Comparative influence of pH ez, R., Lavari, L., Audero, G., Cuatrin, A., Zaritzky, N., Reinheimer, J., et al. (2013).
Pa
and heat on whey protein isolate in protecting Lactobacillus plantarum A17 Study of the effects of spray-drying on the functionality of probiotic lactobacilli.
during spray drying. Food Hydrocolloids, 54(Part A), 162e169. https://doi.org/10. International Journal of Dairy Technology, 66(2), 155e161. https://doi.org/10.1111/
1016/j.foodhyd.2015.09.029. 1471-0307.12038.
Khem, S., Woo, M. W., Small, D. M., Chen, X. D., & May, B. K. (2015). Agent selection Passot, S., Cenard, S., Douania, I., Tre lea, I. C., & Fonseca, F. (2012). Critical water
and protective effects during single droplet drying of bacteria. Food Chemistry, activity and amorphous state for optimal preservation of lyophilised lactic acid
166, 206e214. https://doi.org/10.1016/j.foodchem.2014.06.010. bacteria. Food Chemistry, 132(4), 1699e1705. https://doi.org/10.1016/j.foodchem.
Kim, B. J., Park, T., Moon, H. C., Park, S.-Y., Hong, D., Ko, E. H., … Choi, I. S. (2014). 2011.06.012.
Cytoprotective alginate/polydopamine core/shell microcapsules in microbial Peighambardoust, S. H., Golshan Tafti, A., & Hesari, J. (2011). Application of spray
encapsulation. Angewandte Chemie International Edition, 53(52), 14443e14446. drying for preservation of lactic acid starter cultures: A review. Trends in Food
https://doi.org/10.1002/anie.201408454. Science & Technology, 22(5), 215e224. https://doi.org/10.1016/j.tifs.2011.01.009.
Kumar, P., & Mishra, H. N. (2004). Yoghurt powder-a review of process technology, Perdana, J., Bereschenko, L., Fox, M. B., Kuperus, J. H., Kleerebezem, M., Boom, R. M.,
storage and utilization. Food and Bioproducts Processing, 82(2), 133e142. https:// et al. (2013). Dehydration and thermal inactivation of Lactobacillus plantarum
doi.org/10.1205/0960308041614918. WCFS1: Comparing single droplet drying to spray and freeze drying. Food
Lavari, L., Ianniello, R., Pa ez, R., Zotta, T., Cuatrin, A., Reinheimer, J., … Vinderola, G. Research International, 54(2), 1351e1359. https://doi.org/10.1016/j.foodres.2013.
(2015). Growth of Lactobacillus rhamnosus 64 in whey permeate and study of 09.043.
the effect of mild stresses on survival to spray drying. LWT - Food Science and Perdana, J., Fox, M. B., Siwei, C., Boom, R. M., & Schutyser, M. A. I. (2014). Interactions
Technology, 63(1), 322e330. https://doi.org/10.1016/j.lwt.2015.03.066. between formulation and spray drying conditions related to survival of Lacto-
Lavari, L., Paez, R., Cuatrin, A., Reinheimer, J., & Vinderola, G. (2014). Use of cheese bacillus plantarum WCFS1. Food Research International, 56, 9e17. https://doi.org/
whey for biomass production and spray drying of probiotic lactobacilli. Journal 10.1016/j.foodres.2013.12.007.
of Dairy Research, 81(03), 267e274. https://doi.org/10.1017/ rez-Chabela, M. L., Lara-Labastida, R., Rodriguez-Huezo, E., & Totosaus, A. (2013).
Pe
S0022029914000156. Effect of spray drying encapsulation of thermotolerant lactic acid bacteria on
Lebeer, S., Vanderleyden, J., & De Keersmaecker, S. C. J. (2008). Genes and molecules meat batters properties. Food and Bioprocess Technology, 6(6), 1505e1515.
of lactobacilli supporting probiotic action. Microbiology and Molecular Biology https://doi.org/10.1007/s11947-012-0865-y.
Reviews, 72(4), 728e764. https://doi.org/10.1128/MMBR.00017-08. Pinto, S. S., Fritzen-Freire, C. B., Benedetti, S., Murakami, F. S., Petrus, J. C. C.,
Lee, B., Tachon, S., Eigenheer, R. A., Phinney, B. S., & Marco, M. L. (2015). Lactobacillus Prude ^ncio, E. S., et al. (2015). Potential use of whey concentrate and prebiotics
casei low-temperature, dairy-associated proteome promotes persistence in the as carrier agents to protect Bifidobacterium-BB-12 microencapsulated by spray
mammalian digestive tract. Journal of Proteome Research, 14(8), 3136e3147. drying. Food Research International, 67, 400e408. https://doi.org/10.1016/j.
https://doi.org/10.1021/acs.jproteome.5b00387. foodres.2014.11.038.
Lee, B., Yin, X., Griffey, S. M., & Marco, M. L. (2015). Attenuation of colitis by Poddar, D., Das, S., Jones, G., Palmer, J., Jameson, G. B., Haverkamp, R. G., et al. (2014).
Lactobacillus casei BL23 is dependent on the dairy delivery matrix. Applied and Stability of probiotic Lactobacillus paracasei during storage as affected by the
Environmental Microbiology, 01360e01415. AEM https://doi.org/10.1128/AEM. drying method. International Dairy Journal, 39(1), 1e7. https://doi.org/10.1016/j.
01360-15. idairyj.2014.04.007.
Leslie, S., Israeli, E., Lighthart, B., Crowe, J., & Crowe, L. (1995). Trehalose and sucrose Rajam, R., & Anandharamakrishnan, C. (2015). Microencapsulation of Lactobacillus
protect both membranes and proteins in intact bacteria during drying. Applied plantarum (MTCC 5422) with fructooligosaccharide as wall material by spray
and Environmental Microbiology, 61(10), 3592e3597. drying. LWT - Food Science and Technology, 60(2, Part 1), 773e780. https://doi.
Leverrier, P., Vissers, J. P. C., Rouault, A., Boyaval, P., & Jan, G. (2004). Mass spec- org/10.1016/j.lwt.2014.09.062.
trometry proteomic analysis of stress adaptation reveals both common and Rajam, R., Karthik, P., Parthasarathi, S., Joseph, G. S., & Anandharamakrishnan, C.
distinct response pathways in Propionibacterium freudenreichii. Archives of (2012). Effect of whey protein e alginate wall systems on survival of micro-
Microbiology, 181(3), 215e230. https://doi.org/10.1007/s00203-003-0646-0. encapsulated Lactobacillus plantarum in simulated gastrointestinal conditions.
Lian, W. C., Hsiao, H. C., & Chou, C. C. (2002). Survival of bifidobacteria after spray- Journal of Functional Foods, 4(4), 891e898. https://doi.org/10.1016/j.jff.2012.06.
drying. International Journal of Food Microbiology, 74(1e2), 79e86. https://doi. 006.
org/10.1016/S0168-1605(01)00733-4. Ranadheera, C. S., Evans, C. A., Adams, M. C., & Baines, S. K. (2015). Microencapsu-
Li, X., Lin, S. X. Q., Chen, X. D., Chen, L., & Pearce, D. (2006). Inactivation kinetics of lation of Lactobacillus acidophilus LA-5, Bifidobacterium animalis subsp. lactis BB-
probiotic bacteria during the drying of single milk droplets. Drying Technology, 12 and Propionibacterium jensenii 702 by spray drying in goat's milk. Small
24(6), 695e701. https://doi.org/10.1080/07373930600684890. Ruminant Research, 123(1), 155e159. http://doi.org/10.1016/j.smallrumres.2014.
Linders, L. J. M., Meerdink, G., & Van ’t Riet, K. (1996). Influence of temperature and 10.012.
drying rate on the dehydration inactivation of Lactobacillus plantarum. Food and rquez, R. (2009). Spray drying of a vaginal probiotic strain
Riveros, B., Ferrer, J., & Bo
Bioproducts Processing: Transactions of the Institution of of Chemical Engineers, C, of Lactobacillus acidophilus. Drying Technology, 27(1), 123e132. https://doi.org/
74(2), 110e114. 10.1080/07373930802566002.
Linders, L. J. M., Meerdink, G., & Van’t Riet, K. (1997). Effect of growth parameters on Romantsov, T., Guan, Z., & Wood, J. M. (2009). Cardiolipin and the osmotic stress
the residual activity of Lactobacillus plantarum after drying. Journal of Applied responses of bacteria. Biochimica et Biophysica Acta (BBA) - Biomembranes,
Microbiology, 82(6), 683e688. 1788(10), 2092e2100. https://doi.org/10.1016/j.bbamem.2009.06.010.
Liu, H., Gong, J., Chabot, D., Miller, S. S., Cui, S. W., Ma, J., … Wang, Q. (2015). Pro- Rossi, F., Zotta, T., Iacumin, L., & Reale, A. (2016). Theoretical insight into the heat
tection of heat-sensitive probiotic bacteria during spray-drying by sodium shock response (HSR) regulation in Lactobacillus casei and L. rhamnosus. Journal
caseinate stabilized fat particles. Food Hydrocolloids, 51, 459e467. https://doi. of Theoretical Biology. https://doi.org/10.1016/j.jtbi.2016.04.029.
org/10.1016/j.foodhyd.2015.05.015. Sabikhi, L., Babu, R., Thompkinson, D. K., & Kapila, S. (2010). Resistance of micro-
Lyu, C., Hu, S., Huang, J., Luo, M., Lu, T., Mei, L., et al. (2016). Contribution of the encapsulated Lactobacillus acidophilus LA1 to processing treatments and
activated catalase to oxidative stress resistance and g-aminobutyric acid pro- simulated gut conditions. Food and Bioprocess Technology, 3(4), 586e593.
duction in Lactobacillus brevis. International Journal of Food Microbiology, 238, https://doi.org/10.1007/s11947-008-0135-1.
S. Huang et al. / Trends in Food Science & Technology 63 (2017) 1e17 17

Sadek, C., Tabuteau, H., Schuck, P., Fallourd, Y., Pradeau, N., Le Floch-Foue re, C., et al. Teixido, N., Can~ ama
s, T. p., Usall, J., Torres, R., Magan, N., & Vin ~ as, I. (2005). Accu-
(2013). Shape, shell, and vacuole formation during the drying of a single mulation of the compatible solutes, glycineebetaine and ectoine, in osmotic
concentrated whey protein droplet. Langmuir, 29(50), 15606e15613. https://doi. stress adaptation and heat shock cross-protection in the biocontrol agent
org/10.1021/la404108v. Pantoea agglomerans CPA-2. Letters in Applied Microbiology, 41(3), 248e252.
Santivarangkna, C., Kulozik, U., & Foerst, P. (2007). Alternative drying processes for https://doi.org/10.1111/j.1472-765X.2005.01757.x.
the industrial preservation of lactic acid starter cultures. Biotechnology Progress, To, B. C. S., & Etzel, M. R. (1997). Spray drying, freeze drying, or freezing of three
23(2), 302e315. different lactic acid bacteria species. Journal of Food Science, 62(3), 576e578.
Santivarangkna, C., Kulozik, U., & Foerst, P. (2008). Inactivation mechanisms of lactic Upadrasta, A., Stanton, C., Hill, C., Fitzgerald, G. F., & Ross, R. P. (2011). Improving the
acid starter cultures preserved by drying processes. Journal of Applied Micro- stress tolerance of probiotic cultures: Recent trends and future directions. In
biology, 105(1), 1e13. E. Tsakalidou, & K. Papadimitriou (Eds.), Stress responses of lactic acid bacteria
Schuck, P., Dolivet, A., Me jean, S., Herve , C., & Jeantet, R. (2013). Spray drying of (pp. 395e438). US: Springer.
dairy bacteria: New opportunities to improve the viability of bacteria powders. Vesterlund, S., Salminen, K., & Salminen, S. (2012). Water activity in dry foods
International Dairy Journal, 31(1), 12e17. https://doi.org/10.1016/j.idairyj.2012. containing live probiotic bacteria should be carefully considered: A case study
01.006. with Lactobacillus rhamnosus GG in flaxseed. International Journal of Food
Schuck, P., Jeantet, R., Bhandari, B., Chen, X. D., Perrone, I. T., Carvalho, A. F. de, et al. Microbiology, 157(2), 319e321. https://doi.org/10.1016/j.ijfoodmicro.2012.05.
(2016). Recent advances in spray drying relevant to the dairy industry: A 016.
comprehensive critical review. Drying Technology, 34(15), 1773e1790. https:// Wang, J., Huang, S., Fu, N., Jeantet, R., & Chen, X. D. (2016). Thermal aggregation of
doi.org/10.1080/07373937.2016.1233114. calcium-fortified skim milk enhances probiotic protection during convective
Schuck, P., Jeantet, R., Tanguy, G., Me jean, S., Gac, A., Lefebvre, T., … Martineau, M. droplet drying. Journal of Agricultural and Food Chemistry. https://doi.org/10.
(2015). Energy consumption in the processing of dairy and feed powders by 1021/acs.jafc.6b02205.
evaporation and drying. Drying Technology, 33, 176e184. https://doi.org/10. Wang, Y. C., Yu, R. C., & Chou, C. C. (2004). Viability of lactic acid bacteria and
1080/07373937.2014.942913. bifidobacteria in fermented soymilk after drying, subsequent rehydration and
Schutyser, M. A. I., Perdana, J., & Boom, R. M. (2012). Single droplet drying for storage. International Journal of Food Microbiology, 93(2), 209e217. https://doi.
optimal spray drying of enzymes and probiotics. Trends in Food Science & org/10.1016/j.ijfoodmicro.2003.12.001.
Technology, 27(2), 73e82. https://doi.org/10.1016/j.tifs.2012.05.006. Wood, J. M. (2011). Bacterial osmoregulation: A paradigm for the study of cellular
Sheehan, V. M., Sleator, R. D., Fitzgerald, G. F., & Hill, C. (2006). Heterologous homeostasis. Annual Review of Microbiology, 65(1), 215e238. https://doi.org/10.
expression of BetL, a betaine uptake system, enhances the stress tolerance of 1146/annurev-micro-090110-102815.
Lactobacillus salivarius UCC118. Applied and Environmental Microbiology, 72(3), Wu, V. C. H. (2008). A review of microbial injury and recovery methods in food. Food
2170e2177. Microbiology, 25(6), 735e744. https://doi.org/10.1016/j.fm.2008.04.011.
Silva, J., Carvalho, A. S., Ferreira, R., Vitorino, R., Amado, F., Würth, R., Ho €rmannsperger, G., Wilke, J., Foerst, P., Haller, D., & Kulozik, U. (2015).
Domingues, P., … Gibbs, P. A. (2005). Effect of the pH of growth on the survival Protective effect of milk protein based microencapsulation on bacterial survival
of Lactobacillus delbrueckii subsp bulgaricus to stress conditions during spray- in simulated gastric juice versus the murine gastrointestinal system. Journal of
drying. Journal of Applied Microbiology, 98(3), 775e782. Functional Foods, 15, 116e125. https://doi.org/10.1016/j.jff.2015.02.046.
Silva, J., Carvalho, A. S., Teixeira, P., & Gibbs, P. A. (2002). Bacteriocin production by Wu, R., Song, X., Liu, Q., Ma, D., Xu, F., Wang, Q., … Wu, J. (2016). Gene expression of
spray-dried lactic acid bacteria. Letters in Applied Microbiology, 34(2), 77e81. Lactobacillus plantarum FS5-5 in response to salt stress. Annals of Microbiology,
Simpson, P. J., Stanton, C., Fitzgerald, G. F., & Ross, R. P. (2005). Intrinsic tolerance of 66(3), 1181e1188. https://doi.org/10.1007/s13213-016-1199-1.
Bifidobacterium species to heat and oxygen and survival following spray drying Wu, R., Zhang, W., Sun, T., Wu, J., Yue, X., Meng, H., et al. (2011). Proteomic analysis
and storage. Journal of Applied Microbiology, 99(3), 493e501. of responses of a new probiotic bacterium Lactobacillus casei Zhang to low acid
Song, S. H., Cho, Y. H., & Park, J. (2003). Microencapsulation of Lactobacillus casei YIT stress. International Journal of Food Microbiology, 147(3), 181e187. https://doi.
9018 using a microporous glass membrane emulsification system. Food Engi- org/10.1016/j.ijfoodmicro.2011.04.003.
neering and Physical Properties, 68(1), 195e200. Yonekura, L., Sun, H., Soukoulis, C., & Fisk, I. (2014). Microencapsulation of Lacto-
Sosa, N., Gerbino, E., Golowczyc, M. A., Schebor, C., Go  mez-Zavaglia, A., & bacillus acidophilus NCIMB 701748 in matrices containing soluble fibre by spray
Tymczyszyn, E. E. (2016). Effect of galacto-oligosaccharides: Maltodextrin drying: Technological characterization, storage stability and survival after
matrices on the recovery of Lactobacillus plantarum after spray-drying. Food in vitro digestion. Journal of Functional Foods, 6, 205e214. https://doi.org/10.
Microbiology, 584. https://doi.org/10.3389/fmicb.2016.00584. 1016/j.jff.2013.10.008.
Soukoulis, C., Behboudi-Jobbehdar, S., Yonekura, L., Parmenter, C., & Fisk, I. (2014). Zhang, L., Huang, S., Ananingsih, V. K., Zhou, W., & Chen, X. D. (2014). A study on
Impact of milk protein type on the viability and storage stability of micro- Bifidobacterium lactis Bb12 viability in bread during baking. Journal of Food
encapsulated Lactobacillus acidophilus NCIMB 701748 using spray drying. Food Engineering, 122, 33e37. https://doi.org/10.1016/j.jfoodeng.2013.08.029.
and Bioprocess Technology, 7(5), 1255e1268. https://doi.org/10.1007/s11947- Zhang, Y., Lin, J., & Zhong, Q. (2016). Effects of media, heat adaptation, and outlet
013-1120-x. temperature on the survival of Lactobacillus salivarius NRRL B-30514 after spray
Strasser, S., Neureiter, M., Geppl, M., Braun, R., & Danner, H. (2009). Influence of drying and subsequent storage. LWT - Food Science and Technology, 74, 441e447.
lyophilization, fluidized bed drying, addition of protectants, and storage on the https://doi.org/10.1016/j.lwt.2016.08.008.
viability of lactic acid bacteria. Journal of Applied Microbiology, 107(1), 167e177. Zhang, Z., Luan, C., Zhang, H., Zhang, L., & Hao, Y. (2015). Effects of spray drying on
Sunny-Roberts, E. O., & Knorr, D. (2009). The protective effect of monosodium Lactobacillus plantarm BM-1 viability, resistance to simulated gastrointestinal
glutamate on survival of Lactobacillus rhamnosus GG and Lactobacillus rham- digestion and storage stability. Drying Technology, 34(2), 177e184. https://doi.
nosus E-97800 (E800) strains during spray-drying and storage in trehalose- org/10.1080/07373937.2015.1021009.
containing powders. International Dairy Journal, 19(4), 209e214. https://doi. Zheng, X., Fu, N., Huang, S., Jeantet, R., & Chen, X. D. (2016). Exploring the protective
org/10.1016/j.idairyj.2008.10.008. effects of calcium-containing carrier against drying-induced cellular injuries of
Teixeira, P. C., Castro, M. H., & Kirby, R. M. (1995). Spray drying as a method for probiotics using single droplet drying technique. Food Research International.
preparing concentrated cultures of Lactobacillus bulgaricus. Journal of Applied https://doi.org/10.1016/j.foodres.2016.10.034.
Bacteriology, 78(4), 456e462. https://doi.org/10.1111/j.1365-2672.1995. Zotta, T., Guidone, A., Ianniello, R. G., Parente, E., & Ricciardi, A. (2013). Temperature
tb03433.x. and respiration affect the growth and stress resistance of Lactobacillus planta-
Teixeira, P. C., Castro, M. H., & Kirby, R. M. (1996). Evidence of membrane lipid rum C17. Journal of Applied Microbiology, 115(3), 848e858. https://doi.org/10.
oxidation of spray-dried Lactobacillus bulgaricus during storage. Letters in 1111/jam.12285.
Applied Microbiology, 22(1), 34e38. Zuo, F., Yu, R., Khaskheli, G. B., Ma, H., Chen, L., Zeng, Z., … Chen, S. (2014). Ho-
Teixeira, P. C., Castro, M. H., Malcata, F. X., & Kirby, R. M. (1995). Survival of Lacto- mologous overexpression of alkyl hydroperoxide reductase subunit C (ahpC)
bacillus delbrueckii ssp. bulgaricus following spray-drying. Journal of Dairy Sci- protects Bifidobacterium longum strain NCC2705 from oxidative stress. Research
ence, 78(5), 1025e1031. https://doi.org/10.3168/jds.S0022-0302(95)76718-2. in Microbiology, 165(7), 581e589. https://doi.org/10.1016/j.resmic.2014.05.040.

You might also like