SlideShare a Scribd company logo
1 of 10
Download to read offline
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
Process Biochemistry xxx (2016) xxx–xxx
Contents lists available at ScienceDirect
Process Biochemistry
journal homepage: www.elsevier.com/locate/procbio
Integrated sophorolipid production and gravity separation
Ben M. Dolman, Candice Kaisermann, Peter J. Martin, James B. Winterburn∗
School of Chemical Engineering and Analytical Science, The Mill, The University of Manchester, Manchester, M13 9PL, UK
a r t i c l e i n f o
Article history:
Received 9 November 2016
Received in revised form
13 December 2016
Accepted 21 December 2016
Available online xxx
Chemical compounds:
Sophorolipid (PubChem CID: 11856871)
Keywords:
Integrated separation
Sophorolipid
Settling
Candida bombicola
Biosurfactant
Glycolipid
a b s t r a c t
A novel method for the integrated gravity separation of sophorolipid from a fermentation broth has been
developed, enabling removal of a sophorolipid phase of either higher or lower density than the bulk fer-
mentation broth, while cells and other media components are recirculated and returned to the bioreactor.
The capability of the separation system to recover an enriched sophorolipid product phase was demon-
strated on three sophorolipid producing fed batch fermentations using Candida bombicola, giving an 11%
reduction in fermenter volume required whilst maintaining sophorolipid production. Sophorolipid recov-
eries of up to 86% (280 g) of the total produced over a whole fermentation were achieved at an enrichment
of up to 9. Furthermore, the broth viscosity reduction achieved by removal of the sophorolipid phase
enabled a 34% reduction in mixing power to maintain the same dissolved oxygen level by the end of the
fermentation, with a 9% average reduction over the course of the fermentation. Fermentation duration
could be extended to 1023 h, allowing production of 623 g sophorolipid from 1 l initial batch volume.
These benefits could lead to a substantial decrease in the cost of sophorolipid production, making high
volume applications such as enhanced oil recovery economically feasible.
© 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
1. Introduction
Sophorolipids are microbially produced glycolipid biosurfac-
tants, which are rapidly increasing their market share of the 27
billion USD global surfactant market [1]. While several yeast strains
are able to synthesize sophorolipids, most research and industrial
use is focused on Candida bombicola ATCC 22214, the organism used
in this study [2]. Sophorolipids consist of a hydrophilic sophorose
disaccharide bound to a hydrophobic fatty acid with a typical chain
length of 16–18 carbon atoms. The fatty acid may be joined by
an ester bond to the second glucose monomer, giving a lactonic
sophorolipid, or joined to only one glucose monomer, giving an
acidic sophorolipid due to the unbound fatty acid. These and other
differences in the fatty acid chain and acetylation of the sophorose
molecules give a range of different structures and properties. Two
common structures representing lactonic and acidic sophorolipids
are shown in Fig. 1 [2].
Sophorolipids are produced industrially by a number of compa-
nies, who often utilize sophorolipids’ detergent and low foaming
properties in a variety of formulated cleaning products [3]. The
therapeutic properties of sophorolipids have allowed them to be
commercialized in anti-dermatitis soap and other body washes,
∗ Corresponding author.
E-mail address: james.winterburn@manchester.ac.uk (J.B. Winterburn).
and in a cream to reduce oily skin by MG Intobio Co and Soliance.
There is ongoing research into potential medical applications of
sophorolipid, with anti-cancer, anti-HIV, antimicrobial and anti-
biofilm activity being investigated [4–6]. Sophorolipids also have
potential for use in low cost, high volume applications such as
bioremediation and enhanced oil recovery if production costs can
be significantly reduced [7,8].
In sophorolipid producing fermentations, product concentra-
tions of over 300 g l−1, with productivities of over 1 g l−1 h−1 are
routinely achieved [9–11]. Sophorolipid producing fermentations
begin with a cell growth phase, which typically lasts until the
nitrogen in the media is depleted, at which point the sophorolipid
production rate increases significantly, if both a hydrophilic and a
hydrophobic carbon source are present [12]. The sophorolipid pro-
duction phase normally lasts for around 200 h, at which point the
dissolved oxygen level in the fermenter cannot be maintained due
to oxygen mass transfer limitation.
This dissolved oxygen reduction is caused by the high viscosity
of the sophorolipid produced, meaning the fermentation must be
stopped and the sophorolipid recovered [12,13]. It is well known
that the presence of a separate sophorolipid phase in the bioreactor
significantly reduces the oxygen mass transfer coefficient, kLa, by
both providing a resistance to mass transfer across the air/liquid
interface and increasing the viscosity of the medium, which results
in oxygen limitation, increased stirring power requirements and
non-homogeneity in the bioreactor [12–14].
http://dx.doi.org/10.1016/j.procbio.2016.12.021
1359-5113/© 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
2 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx
Fig. 1. Molecular structure of common lactonic (left) and acidic (right) sophorolipids. A lactonic bond can be seen joining the fatty acid chain to the second glucose monomer
of the sophorose in the lactonic sophorolipid, where the acidic sophorolipid has a free carboxylic acid to end its fatty acid chain.
The physical form of sophorolipids is dependent on the con-
ditions under which they are produced, which directly affect
the proportions of acidic and lactonic sophorolipids produced.
Sophorolipids typically separate from the fermentation broth as
a crystalline material if the lactonic to acidic ratio is high and the
hydrophobic carbon source concentration is low. The sophorolipids
otherwise form a viscous second phase of around 50% sophorolipid
and 50% water, which may sit below residual oil at the surface of
the broth or sink to the bottom of the bioreactor when agitation is
stopped [10,15,16].
These properties are commonly exploited at the end of a fermen-
tation to give an easy, crude separation of the sophorolipid from the
fermentation broth, either by crystal decantation, crystal filtration
or decantation of the sophorolipid gel [11,16,17]. These techniques
have not previously been used effectively to recover sophorolipids
during fermentation.
Industrially, there are a number of costs associated with
repeated batch cycles for sophorolipid fermentation, in terms of
downtime between cycles, cleaning costs and the lengthy inocu-
lum preparation required for large scale production [18,19]. There
are numerous proposed partial solutions to this problem in the lit-
erature. For example, a portion of the broth can be removed and
replaced with fresh media, allowing high productivity to be main-
tained for seven 80–130 h cycles, nevertheless removing biomass
proportionally to other components [19]. Sophorolipid settling by
gravity within a fermentation vessel or shake flask has previously
been demonstrated for small scale sophorolipid production. Signif-
icant benefits of sophorolipid separation have been shown, with a
doubling of the duration of sophorolipid production and little effect
on production after 15 min without agitation or aeration demon-
strated by Guilmanov et al. [20], and a productivity increase from
1.38 to 1.89 g l−1 h−1 shown by Marchal et al. [19]. Both studies rely
on gravity settling within the fermentation vessel or shake flask,
however, making scale up impractical due to the excessive settling
distances present if this technique were applied at industrial scale.
Effective integrated separation techniques have been developed
for other biosurfactant systems, notably foam fractionation for
hydrophobin proteins, surfactin and rhamnolipids, but there have
been no successful scalable attempts at integrated separation for
sophorolipid production [21–23].
This paper details a novel technique, based on an integrated
gravity settling column, for removing the sophorolipid phase from
the fermentation broth during fermentation, reducing the fermen-
tation volume required which allows continued substrate feeding
and provides a concentrated product phase. We also demonstrate,
for the first time, the application of this technique to extend the
production phase of a fermentation beyond 1000 h, significantly
increasing batch production.
2. Methodology
2.1. Fermentation
Sophorolipid was produced by fed batch fermentation using C.
bombicola ATCC 22214 in an Electrolab Fermac 320 fermentation
system (Electrolab, UK) with a 2 l maximum working volume, H:D
of 2, two 55 mm diameter 6-bladed Rushton-type impellers and an
initial working volume of 1 l, with a stirring rate of 200–800 rpm.
This was sufficient to disperse vegetable oil within the bioreactor.
Growth medium for the fermentations, preculture and agar plates
contained 6 g l−1 yeast extract and 5 g l−1 peptone. The initial con-
centration of glucose in all fermentations preculture and agar plates
was 100 g l−1, with an initial rapeseed oil concentration of 50 g l−1
in the fermenters, 100 g l−1 in the preculture and 0 g l−1 in the agar
plates.
Four fermentations were carried out. Fermentation 1 was con-
ducted in a conventional manner without separation. Fermentation
2 was directly comparable to fermentation 1 except that the in situ
separator was used to remove product from the top of the separa-
tor. Fermentation 3 was controlled to give a product that separated
from the bottom of the separator. Fermentation 4 was carried out
for an extended period of time to demonstrate the potential to
utilise the integrated separation to extend the fermentation period
and give higher batch production, and controlled to give separation
from the surface of the broth. C. bombicola was first transferred from
cryogenic storage (−80 ◦C) onto agar plates, and incubated at 25 ◦C
for 48 h. Single colonies from these plates were then used to inocu-
late 50 ml of medium in 250 ml shake flasks, which were incubated
at 25 ◦C and 200 rpm for 30 h. This inoculum was diluted to an opti-
cal density of 20 at 600 nm with fresh media and 100 ml used to
inoculate the fermenter.
Fermentations were run at 25 ◦C, and dissolved oxygen was con-
trolled to 30% by varying the stirrer speed, whilst maintaining a
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 3
constant aeration rate of 1 l min−1. Fermenter pH was controlled to
a value of 3.5 by the addition of 3 M sodium hydroxide.
The feeding rates of rapeseed oil and glucose were similar for
fermentation one and two, to facilitate comparison, with different
feeding rates used for fermentation three and four, to give separa-
tion of sophorolipid to the top and bottom of the fermenter. Feeding
rates of oil were modified during the experiments to maintain a
low concentration, without limiting production, according to the
oil concentration in the samples. Glucose concentration was used
to control the relative density of the sophorolipid phase and the
bulk media to enable effective separation from either the top or
bottom of the separator, as well as being an important substrate
for sophorolipid production. The feed profiles are shown in Fig. 4.
The total sophorolipid produced was calculated by adding the
mass of sophorolipid in the fermenter and the mass of sophorolipid
removed from the fermenter using the separator. Contamination
was tested for visually using microscopy and by streak plating.
2.2. Separation
Integrated separation of the sophorolipid from the fermentation
broth was carried out using an in house built settling column, as
shown in Fig. 2. The integrated arrangement of the bioreactor and
settling column is shown in Fig. 3 with the settling column sup-
ported at an angle of 30◦ from horizontal. The separator was rinsed
with 70% ethanol before attaching to the fermenter. Sophorolipid
separation was carried out intermittently, based on visual observa-
tion of a sample taken from the bioreactor. When a significant layer
of sophorolipid rich phase could be seen at either the top or bot-
tom of the sample in the universal bottle within 2 min separation
was initiated. Prior experiments indicated that at these conditions
separation is effective.
During separation, which was used intermittently during fer-
mentation, sophorolipid rich fermentation broth was continuously
circulated from the fermenter, through the settling column and
back to the fermenter. This was pumped from the fermenter
through a stainless steel tube with an inlet 20 mm from the bot-
tom of the fermenter, and then in 8 mm external diameter silicon
tubing of 1 mm wall thickness to and from the separator. The flow
rate of media into and out of the settler was controlled to 1 ml s−1
using Matson Marlow 502 S and 503 U pumps (Watson Marlow,
UK). This flowrate was based on the results of preliminary experi-
ments, giving a residence time in the settling column of 76 s, with
a total residence time in the column and tubing of 137 s. In the set-
tling column, the sophorolipid phase separates out towards either
the top or bottom of the column, depending on the relative density
of the sophorolipid and bulk media. Initially, broth is continuously
circulated and the sophorolipid product collects in the settling col-
umn. When the sophorolipid phase accumulating in the separator
reached 50% of the height of the separator, which typically occurred
after around three minutes of separator operation, the outlet pump
was started to continuously remove the sophorolipid product phase
at a rate controlled between 0.5 and 2 ml min−1, depending on the
accumulation or reduction of the sophorolipid phase in the set-
tling vessel. The separation was stopped when the separation rate
dropped below 0.5 ml min−1, until the condition for separation was
again observed.
2.3. Hydrodynamics
To determine the effect of the sophorolipid phase on the agita-
tion requirements, the sophorolipid enriched fractions, which had
been separated from the bioreactor using the separator over the
course of fermentation 2, were pooled and returned to the fer-
menter at 308 h after the start of the fermentation over a period of
12 min. The stirrer speed and dissolved oxygen percentages were
then monitored as the fermentation control returned the dissolved
oxygen percentage to the set point. The equation for power number
is shown in Eq. (1);
P = Np n3
D5
(1)
where P is power, Np is power number, is density, n is stirrer speed
and D is stirrer diameter.
Due to identical fermenters being used and assuming the den-
sity is constant (as density changes during the fermentation are
relatively small), the power input as a function of power number
can be calculated during the fermentation. The power input was
integrated over time to determine the power input for the whole
fermentation, with the power input for fermentations equated until
the time point of the first separation.
Separation performance was measured in terms of enrichment
and recovery, which are defined in Eqs. (2) and (3);
enrichment =
Cp
Cf
(2)
recovery (%) =
Cp × Vp
Cf × Vf
× 100 (3)
where Cp is the sophorolipid concentration in the product, Cf is the
product concentration in the fermenter before separation, Vp is the
volume of product phase recovered, and Vf is the initial volume of
liquid in the fermenter.
2.4. Analytical techniques
For all analysis, 5 ml of broth was removed from the bioreactor or
5 ml product was taken from the sophorolipid collection vessel con-
nected to the separator. The sample was centrifuged at 5000 rpm
for 5 min using a Sigma 6–16S centrifuge (Sigma laboratory cen-
trifuges, Germany) and the glucose in the supernatant quantified
using a TrueResult
®
blood glucose monitor (Nipro, Japan).
A hexane extraction to extract residual rapeseed oil fol-
lowed by a triple ethyl acetate extraction to extract sophorolipid
were then applied to the whole sample, with oil concentra-
tion measured gravimetrically from the hexane extraction, and
sophorolipid measured gravimetrically from the pooled ethyl
acetate extracts[24–26]. These extracts were dried to constant
weight in weighing dishes at ambient temperature for 30 h.
Cell growth was determined by both dry cell weight and optical
density measurement. After the aforementioned hexane and ethyl
acetate extractions, 8 ml distilled water was added to the remainder
of the sample in the centrifuge tubes, which were then centrifuged
at 8000 rpm for 10 min. The supernatant was discarded and the
resulting cell pellet was resuspended in 8 ml distilled water. This
cell suspension was transferred to drying trays, which were dried
to constant weight at 90 ◦C in a drying oven. Optical density was
used as a proxy for dry cell weight when diluting the inoculum, at
a wavelength of 600 nm.
The structure of the sophorolipids produced was determined
with negative electrospray ionisation mass spectrometry, using
an Agilent 6520 QTOF mass spectrometer (Agilent, United States).
Samples were prepared by redissolving the ethyl acetate extracts,
i.e. sophorolipids, in ethyl acetate, and filtering using a 0.2 ␮m filter.
Flow injection analysis was used, at 0.3 ml min−1, 50% acetonitrile,
0.1% formic acid and 49.9 % water, with an injection volume of 2 ␮l.
The viscosity of the product phase was measured using an
AR2000 controlled rotational rheometer with cone geometry (TA
Instruments, USA).
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
4 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx
Fig. 2. Diagram of custom built sophorolipid separator used for this study. Plan view, side view and end view are shown, all dimensions are internal dimensions shown in
mm..
3. Results and discussion
A novel gravity sophorolipid separation technique was success-
fully applied to three fermentations. Results from one fermentation
without separation, fermentation 1, are presented alongside results
from three fermentations with separation, fermentations 2, 3 and
4. Fermentations 1 and 2, without and with integrated separa-
tion respectively, are directly comparable, to allow evaluation of
the effect of separation, but due to feeding rate control to enable
sophorolipid to be recovered from the bottom of the separator
fermentation 3 is not directly comparable. Fermentation 4 was
intended to demonstrate an extended fermentation, and so is also
not directly comparable. The separation was run periodically in
all fermentations, when sufficient sophorolipid phase had accu-
mulated, with the majority of the available sophorolipid phase
separated.
3.1. Fermentations
Fig. 4 shows the feeding rate of substrates for the fermentations
presented, which enabled control of the sophorolipid phase to sep-
arate from the surface or bottom of the separator, whilst also being
an important parameter for sophorolipid production. The progress
of the fermentations over time is presented in Fig. 5, and the key
metrics from these fermentations are presented in Table 1.
Fig. 5 shows the progress of fermentation 1, without separation,
and fermentations 2, 3 and 4, during which sophorolipid product
was separated from the fermentation broth. In fermentation 2 and
4, sophorolipid was recovered at the top of the integrated gravity
separator, and in fermentation 3 the sophorolipid was collected
from the bottom of the separator. In fermentation 2, separation
was carried out at 111, 184 and 261 h, in fermentation 3 at 72, 281,
355 and 376 h and in fermentation 4 at 86, 111, 160, 186, 232 and
540 h. No separation of the sophorolipid phase was carried out in
fermentation 1.
In fermentation 2, the glucose concentration initially rose, and
remained above 50 g l−1 for the majority of the fermentation, which
led to the sophorolipid rising to the surface of the fermentation
broth without agitation. A high glucose concentration throughout
Table 1
Key metrics for all sophorolipid producing fermentations. Fermentation 2, 3 and
4 used sophorolipid separation, with fermentation 1 included for comparison. All
metrics are the result of unique fermentations. Due to volume changes caused by
substrate addition and product removal, productivity is based on the initial fer-
menter working volume and the total sophorolipid produced.
Fermentation
1 2 3 4
Product separation none top bottom top
Duration (h) 305 305 379 1023
Yield substrate consumed (g g−1
) 0.43 0.53 0.42 0.53
Yield substrate fed (g g−1
) 0.33 0.37 0.39 0.47
Productivity (g l−1
h−1
) 1.07 1.07 0.71 0.61
Maximum fermenter volume (l) 1720 1544 1350 1550
Total sophorolipid produced (g) 325 325 270 623
Total dry cell weight (g) 16.1 21.0 16.5 32.1
Yield product on biomass (g g−1
) 20.2 15.5 16.4 19.4
fermentation 4 meant the sophorolipid was also separated from the
top of the separator during this experiment.
Sophorolipid was first separated from the bottom of the sepa-
rator at 71.5 h in fermentation 3, when a sophorolipid phase could
be observed to settle in a sample bottle within 2 min. Settling was
not possible after this until 283 h due to the high residual glucose
concentrations caused by pulse glucose feeding, which was used to
ensure good sophorolipid production. Whilst the relative density of
the sophorolipid phase and the broth also depend on other factors,
a glucose concentration of 50 g l−1 tends to represent a threshold
of sophorolipid phase separation to the surface or the bottom of
the separator. This is because higher glucose concentrations lead
to higher media densities, meaning the sophorolipid phase is rela-
tively less dense the higher the glucose concentration. After 283 h
the glucose concentration had dropped sufficiently for settling to
be used again. Lower glucose feeding could enable sophorolipid
settling throughout the fermentation, though this might impact
sophorolipid production.
Fermentations 1 and 2, which were identical apart from the
application of integrated sophorolipid separation in fermentation
2, each produced 325 g sophorolipid, with 270 g sophorolipid pro-
duced during fermentation 3. The dry cell weight production, of
16–21 g, and the yields of product on substrate, of 0.33-0.39 are
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 5
Fig. 3. Integrated fermentation system. Bioreactor is shown on the left, with the separator in the center, and the product collection vessel on the right. Broth is pumped
from the bioreactor into the separator, and recirculated back to the bioreactor. Product is pumped from the separator into the collection vessel. This can be used for; (a) −
sophorolipid phase density higher than fermentation broth. (b)− sophorolipid phase density lower than fermentation broth.
in line with other results in the literature, as are the values for
total sophorolipid produced [9,25]. The use of the separator appears
to have little effect on the total production of sophorolipids over
the same time period, with identical values recorded for the two
comparable fermentations.
In fermentation 2 with separation, 21 g of biomass were pro-
duced, more than the 16 g of cell biomass produced in fermentation
1, with product yields on biomass of 15.5 g g−1 and 20.2 g g−1
respectively. The reason for the reduction in product yield on
biomass in fermentations 2 and 3 is not known, but with further
optimisation fermentation 4, with separation, reached a product
yield on biomass of 19.4 g g−1.
The highest working volume reached during a fermentation dic-
tates the overall fermenter volume required, and the high feeding
rates used during sophorolipid producing fermentations lead to a
large increase in the working volume required over time, much of
which is only required in the later stages of the fermentation.
The use of integrated sophorolipid separation in fermentation
2 decreased the fermenter working volume required by removing
523 ml broth from the fermenter using the separator. This meant
only 1540 ml working volume was needed for fermentation 2 com-
pared to 1720 ml in fermentation 1 without separation. This 11%
decrease in volume requirement could reduce bioreactor capital
costs. The corresponding maximum volume for fermentation 3 was
1350 ml, but was not directly comparable due to differences in
feeding rates between fermentations 2 and 3.
The overall productivity of fermentations 1 and 2 were 1.07 g
l−1 h−1 calculated at the starting volume, with a correspond-
ing productivity of 0.77 g l−1 h−1 for fermentation 3. The rate of
sophorolipid production slowed dramatically when the oil was
depleted after around 80 h in fermentations 1–3, reducing from 2 g
l−1 h−1 to 0.6 g l−1 h−1 and increasing the oil feed rate did not return
the productivity to the previous level. Many studies have demon-
strated a fairly constant production rate throughout a fermentation
until the point at which the fermentation had to be stopped due to
dissolved oxygen limitation, and with improved feeding control in
fermentation 4, a productivity of 2.0 g l−1 h−1 was maintained until
158 h [17] [12].
In fermentation 4, fermentation was continued past the point
at which fermentations usually have to be stopped due to product
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
6 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx
Fig. 4. Feeding profiles of glucose and rapeseed oil for all fermentations in this study. Glucose (blue), rapeseed oil (green) and total (red) shown for fermentation 1 (solid, a)
fermentation 2 (dotted, a), fermentation 3 (b) and fermentation 4 (c). (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)
accumulation, and run for a total of 1023 h. This enabled the produc-
tion of 623 g sophorolipid from a 1 l initial broth, which compares
favourably to the highest previous reported titers of around 400 g
l−1, and clearly demonstrates the capacity of integrated separation
to extend the time period of sophorolipid production in fermenta-
tion. This could lead to a dramatic improvement in overall process
productivity, by reducing the proportion of time spent in inoculum
preparation, biomass production and cleaning.
3.2. Separation
The separation results achieved in fermentations 2, 3 and 4 are
shown in Table 2.
During fermentations 2, 3 and 4 with integrated separation, the
majority of the sophorolipid was removed from the fermentation
broth, with 86% of the total sophorolipids produced separated in
fermentation 2, 74% separated in fermentation 3 and 65% separated
in fermentation 4.
Almost no cells and only 8 g of oil were removed by the separa-
tion over the course of fermentation 3, determined by gravimetric
analysis as for fermentation samples. This is because the rate of
settling of the cells was much slower than the settling of the
sophorolipid product, and oil rose to the surface of the separa-
tor rather than sinking to the bottom of the separator with the
sophorolipid. Cell removal was also negligible in fermentation 2,
though 68 g of oil was removed, which was reduced by better feed-
ing rate control in fermentation 4, where only 20 g oil was removed,
while again separating from the surface of the separator. 2.6 g cells
were removed during fermentation 4, which represents a small
proportion of the total biomass, 32.1 g.
The enrichment varied significantly between separations at
different time points, from 2.5 to 9. This is largely due to the
sophorolipid concentration present in the fermenter before the
separation, as there was little variation of the concentration in the
sophorolipid enriched product fraction, of approximately 550 g l−1.
In fermentation 2, when the sophorolipid phase was separated from
the top of the settler, a total of 68 g oil was recovered along with
the sophorolipid during the fermentation, which was the primary
reason for the variations in the concentration of the sophorolipid
phase. There is little scope to improve the product phase concentra-
tion above that demonstrated in fermentation 3 using the current
technique, however, with an increased fermenter volume, the sys-
tem could operate at lower initial sophorolipid concentrations and
so give an improved enrichment.
Sophorolipid recoveries would be expected to improve sig-
nificantly as fermentation scale is increased; in laboratory scale
experiments the separation had to be stopped as the layer of
sophorolipids at the bottom/top of the settling column became too
low, to prevent the media and cell phase being entrained in the
product stream. This minimum sophorolipid phase depth, which
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 7
Fig. 5. Time course of fermentations using sophorolipid separation. Results for fermentation 1 (a), fermentation 2 (b), fermentation 3 (c) and fermentation 4 (d) are presented.
Dry cell (black squares) glucose (blue triangles) rapeseed oil (green circles) and sophorolipid (filled red circles) concentrations are shown. Arrows show sophorolipid separation,
total sophorolipid produced shown by open red circles. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)
Table 2
Separation results for fermentation 2, fermentation 3 and fermentation 4. All metrics are the result of unique fermentations, with the application of the novel integrated
gravity separation developed in this study.
Time (h) Sophorolipid
recovered
(g)
Sophorolipid
concentration
(g l−1
)
Total sophorolipid
present
(g)
Sophorolipid product
concentration
(g l−1
)
Enrichment Recovery at time
point
(%)
Fermentation 2
111 97.1 147.7 259.5 550.6 3.73 37
184 99.2 118.0 165.2 461.6 3.91 60
261 83.8 89.2 96.4 540.9 6.07 87
Total 280.1 86
Total oil removed by separation (g) 8 Total cells removed by separation (g) below detection limit
Fermentation 3
71.5 16.8 103.7 128.4 582.9 5.62 13
281 79.5 168.3 238.2 654.1 3.89 33
355 59.2 109.2 175.3 616.9 5.66 34
376 45.5 106.3 148.9 638.7 6.01 31
Total 201.0 74
Total oil removed by separation (g) 68 Total cells removed by separation (g) below detection limit
Fermentation 4
85.6 93.1 152.7 229.1 443.5 2.90 40
111.3 53.5 97.1 127.8 504.8 5.20 42
159.7 61.9 129.6 186.8 538.1 4.15 33
185.8 57.9 72.5 105.3 567.8 7.83 55
231.5 48.9 51.7 68.3 465.7 9.01 72
540.3 89.0 124.7 191.3 312.2 2.5 47
total 404.3 65
Total oil removed by separation (g) 20 Total cells removed by separation (g) 2.6
must be recycled back to the fermenter, would be identical irre-
spective of fermenter volume while maintaining a given size of
separator, hence a larger total fermenter volume would lead to a
large reduction in residual sophorolipid concentration.
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
8 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx
Fig. 6. The effect of sophorolipid separation on agitation requirements. Dissolved oxygen and stirrer speed profiles showing the increase in stirrer speed required to maintain
the dissolved oxygen concentration after sophorolipid rich fractions separated during fermentation 2 returned to fermenter in fermentation 2 at 308 h.
Whilst the separator was designed for continuous operation
hydrodynamic considerations, in particular the turbulence caused
by inlet and outlet disturbances which become more significant
with decreasing scale, mean it could not be scaled down further to
match the sophorolipid production rates achieved in the 1 l initial
working volume fermenter. At the scale presented in this paper,
the separation occurred at a rate of around 2 ml min−1, or around
1 g min−1 sophorolipid, 30–150 times greater than the production
rate. This separation rate makes it suitable for use with a separa-
tor of 30–60 l volume for continuous separation of the sophorolipid
produced. Product recovery rate is expected to scale proportionally
to the volume of the separating column, so for a new settler design
volume can be increased while maintaining or reducing the ratio of
inertial to viscous forces, i.e. the Reynolds number. Residence time
should be increased proportionally to diameter increase, to allow
the same quantity of sophorolipid to settle or float to the surface at
the same settling/rising rate. The system could easily be connected
for steam in place sterilization at industrial scale.
This is the first study to present the design and demonstrate
the feasibility of a separation system for sophorolipid production
that could be continuously applied, having been shown in this
manuscript to separate sophorolipid whilst production continues
in the bioreactor for periods of more than one hour. It is also the
first integrated separation system applicable to large scale fermen-
tation, because it does not rely on separation within the bioreactor,
and therefore the first to enable an extended sophorolipid produc-
tion period at scale. The reduced bioreactor volume, and reduced
start up and cleaning costs of this system could significantly
improve the economics of sophorolipid production by increasing
the total sophorolipid produced per batch. This would make bulk
application, such as for enhanced oil recovery, a more realistic
proposition.
Other glycolipid biosurfactants, notably rhamnolipids and man-
nosylerythritol lipids (MELs), as well as many other bioproducts
including those used as biofuels, may also form a separate, insol-
uble, phase in a fermentation broth, and so the gravity separation
technique presented in this paper could likely also be applied to
these systems [27,28].
3.3. Effect on hydrodynamics and mass transfer
Fig. 6 shows the dissolved oxygen level and stirrer speed at the
end of fermentation 2, capturing the addition of the sophorolipid
rich fractions which were removed by separation during fermen-
tation 2 and subsequently pooled, and added to the bioreactor at
308 h. The presence of this sophorolipid phase effectively reduced
the volumetric oxygen transfer coefficient, kLa, in the fermenter,
due to its high viscosity of around 0.5 Pa.s, resulting in an increase
in stirrer speed to maintain the dissolved oxygen at the set point.
A stirring rate increase of around 75 rpm, from 500 rpm to 575 rpm
was required to maintain the desired dissolved oxygen level when
the sophorolipid product separated over the whole fermentation
Fig. 7. Mass spectrum of sophorolipid from the end of fermentation 2. Key peaks of C18:1 diacylated acidic sophorolipid at m/z 705 and C18:1 lactonic sophorolipid at m/z
687.
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 9
was added. Given the high viscosity of the sophorolipid phase, it
is possible that turbulence was not attained even at the higher
agitation rate; in this instance a still higher agitation rate would
be required for proper dissolved oxygen control throughout the
bioreactor, leading to larger savings than calculated.
The mixing Power number in the turbulent regime is typically
constant, so changes in impeller speed have a cubic impact on the
mixing power. Removing the sophorolipid phase resulted in a 13%
decrease in impeller speed and a 34% decrease in mixing power
requirement by the end of the fermentation.
The relative power requirements over fermentation 1 and 2
were also compared, with an 18% reduction in power input from the
time point of the first separation observed, and a 9% improvement
when the whole fermentation is taken into account.
There would be some increased power consumption from the
pumping of the fermentation broth between the separator and the
fermenter, but it is expected this would be significantly smaller
than the agitation power at scale, as relatively low pumping flow
rates are used. If these power reductions can be achieved at indus-
trial scale, they can give significant cost savings.
3.4. Sophorolipid structure
Mass spectra of sophorolipid samples were taken to determine
the sophorolipid structures produced during the fermentation,
with the mass spectrum of the sophorolipids taken from the end
of fermentation 2 shown in Fig. 7. The main peak at m/z=705
represents a diacylated acidic C18:1 sophorolipid, with the peak
at 687 representing a diacylated C18:1 lactonic sophorolipid [29].
There were some differences in the ratios of peak heights between
sophorolipid taken from the settler and sophorolipid taken from
the fermentation broth immediately before, suggesting this tech-
nology may have potential to act as a crude separator of different
sophorolipid forms.
Recent research has revealed the enzyme responsible for lac-
tonisation of acidic sophorolipids, enabling the use of genetic
engineering of the genome of Starmerella bombicola, and robust
production and purification techniques to yield a 98% pure lactonic
or acidic sophorolipid, making the application of sophorolipid in
medicinal or personal care products much more likely [3]. Comple-
mentary to the advances in genetic engineering and purification,
this work makes significant progress in solving some of the complex
engineering challenges involved with sophorolipid production, giv-
ing potentially dramatic reductions in production cost and making
large scale application of sophorolipid more feasible.
4. Conclusions
A novel method for the integrated separation of sophorolipid
from a fermentation process has been developed. The design of the
system overcomes the production and processing difficulties asso-
ciated with in situ (i.e. in the fermenter vessel) gravity separation
of sophorolipids for scale up, and with a separator residence time
of less than two minutes the process seemed to have no impact on
further sophorolipid production by the cells. A sophorolipid phase
can be removed from the fermentation broth if the product phase
had higher or lower density than the media, with enrichments of
up to 9, an overall recovery of 86%, and up to 404 g of sophorolipid
recovered from the fermentation broth.
We demonstrated an 11% decrease in bioreactor working vol-
ume requirement when using the separator, due to the removal of
the sophorolipid product phase. Optimised feeding rates and settler
usage could further reduce the volume requirement. By using the
separator, the fermentation could be run for 1023 h, and produce
623 g sophorolipid from a 1 l initial batch, reducing the number of
fermentations required for a given product mass.
An 18% average reduction in stirrer power was demonstrated
over the course of a fermentation once sophorolipid separation was
initiated, which translates to around 9% over the entire fermenta-
tion, with a 34% decrease in power input shown by the end of the
fermentation.
The integrated separation system presented in this paper has
been developed for sophorolipid separation, but could equally be
applied to the production of other insoluble bioproducts, in par-
ticular mannosylerythritol lipids. With correct scaling up, it is
anticipated that the advantages this system offers will lead to a
dramatic improvement of the economics of sophorolipid produc-
tion.
Acknowledgements
Sara Bages Estopa is acknowledged for her invaluable comments
on the paper, and Reynard Spiess and Shaun Leivers are acknowl-
edged for their help with mass spectrometry.
The technology described in this manuscript has been filed for a
patent entitled ‘Improvements in and related to lipid production’.
The authors are grateful for financial support from the UK Engi-
neering and Physical Sciences Research Council (EP/I024905) and
the EPSRC DTA fund, which enabled this work to be conducted.
References
[1] R. Geys, W. Soetaert, I. Van Bogaert, Biotechnological opportunities in
biosurfactant production, Curr. Opin. Biotechnol. 30 (2014) 66–72.
[2] I. Van Bogaert, K. Saerens, C. De Muynck, D. Develter, W. Soetaert, E.J.
Vandamme, Microbial production and application of sophorolipids, Appl.
Microbiol. Biotechnol. 76 (2007) 23–34.
[3] S.L.K.W. Roelants, K. Ciesielska, S.L. De Maeseneire, H. Moens, B. Everaert, S.
Verweire, Q. Denon, B. Vanlerberghe, I. Van Bogaert, P. Van der Meeren, B.
Devreese, W. Soetaert, Towards the industrialization of new biosurfactants:
biotechnological opportunities for the lactone esterase gene from Starmerella
bombicola, Biotechnol. Bioeng. 113 (2016) 550–559.
[4] M.A. Díaz De Rienzo, I.M. Banat, B. Dolman, J. Winterburn, P.J. Martin,
Sophorolipid biosurfactants: possible uses as antibacterial and antibiofilm
agent, N. Biotechnol. 32 (2015) 720–726.
[5] V. Shah, G.F. Doncel, T. Seyoum, K.M. Eaton, I. Zalenskaya, R. Hagver, A. Azim,
R. Gross, Sophorolipids, microbial glycolipids with anti-human
immunodeficiency virus and sperm-immobilizing activities, Antimicrob.
Agents Chemother. 49 (2005) 4093–4100.
[6] L. Shao, X. Song, X. Ma, H. Li, Y. Qu, Bioactivities of sophorolipid with different
structures against human esophageal cancer cells, J. Surg. Res. 173 (2012)
286–291.
[7] A.E. Elshafie, S.J. Joshi, Y.M. Al-Wahaibi, A.S. Al-Bemani, S.N. Al-Bahry, D.
Al-Maqbali, I.M. Banat, Sophorolipids production by Candida bombicola ATCC
22214 and its potential application in microbial enhanced oil recovery, Front.
Microbiol. 6 (2015).
[8] L. Ławniczak, R. Marecik, L. Chrzanowski, Contributions of biosurfactants to
natural or induced bioremediation, Appl. Microbiol. Biotechnol. 97 (2013)
2327–2339.
[9] A.M. Davila, R. Marchal, J.P. Vandecasteele, Sophorose lipid fermentation with
differentiated substrate supply for growth and production phases, Appl.
Microbiol. Biotechnol. 47 (1997) 496–501.
[10] Institut Francais Du Petrole. Method of production of sophorosides by
fermentation with fed batch supply of fatty acid esters or oils. US. Patent. US
5616479A.
[11] U. Rau, S. Hammen, R. Heckmann, V. Wray, S. Lang, Sophorolipids A source for
novel compounds, Ind. Crops Prod. 13 (2001) 85–92.
[12] A.M. Davila, R. Marchal, J.P. Vandecasteele, Kinetics and balance of a
fermentation free from product inhibition: sophorose lipid production by
Candida bombicola, Appl. Microbiol. Biotechnol. 38 (1992) 6–11.
[13] N. Kosaric, F.V. Sukan, Biosurfactants:Production and Utilization-Processes,
Technologies and Economics, CRC Press, Boka Raton, 2014.
[14] U. Rau, C. Manzke, F. Wagner, Influence of substrate supply on the production
of sophorose lipids by Candida bombicola ATCC 22214, Biotechnol. Lett. 18
(1996) 149–154.
[15] H.J. Daniel, R.T. Otto, M. Reuss, C. Syldatk, Sophorolipid production with high
yields on whey concentrate and rapeseed oil without consumption of lactose,
Biotechnol. Lett. 20 (1998) 805–807.
[16] O. Stüwer, R. Hommel, D. Haferburg, H.P. Kleber, Production of crystalline
surface-active glycolipids by a strain of torulopsis apicola, J. Biotechnol. 6
(1987) 259–269.
Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem
(2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021
ARTICLE IN PRESSG Model
PRBI-10893; No.of Pages10
10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx
[17] H.J. Daniel, M. Reuss, C. Syldatk, Production of sophorolipids in high
concentration from deproteinized whey and rapeseed oil in a two stage fed
batch process using Candida bombicola ATCC 22214 and Cryptococcus
curvatus ATCC 20509, Biotechnol. Lett. 20 (1998) 1153–1156.
[18] G.L. Kleman, W.R. Strohl, High cell density and high-productivity microbial
fermentation, Curr. Opin. Biotechnol. 3 (1992) 93–98.
[19] Institut Francals du Petrole. Process for the production of sophorolipids by
cyclic fermentation with feed of fatty acid esters or oils. US. Patent. US
5879913A.
[20] V. Guilmanov, A. Ballistreri, G. Impallomeni, R.A. Gross, Oxygen transfer rate
and sophorose lipid production by Candida bombicola, Biotechnol. Bioeng. 77
(2002) 489–494.
[21] S. Alonso, P.J. Martin, Impact of foaming on surfactin production by Bacillus
subtilis: implications on the development of integrated in situ foam
fractionation removal systems, Biochem. Eng. J. 110 (2016) 125–133.
[22] J. Beuker, A. Steier, A. Wittgens, F. Rosenau, M. Henkel, R. Hausmann,
Integrated foam fractionation for heterologous rhamnolipid production with
recombinant Pseudomonas putida in a bioreactor, AMB Express 6 (2016) 11.
[23] J.B. Winterburn, A.B. Russell, P.J. Martin, Integrated recirculating foam
fractionation for the continuous recovery of biosurfactant from fermenters,
Biochem. Eng. J. 54 (2011) 132–139.
[24] A. Daverey, K. Pakshirajan, Production, characterization, and properties of
sophorolipids from the yeast Candida bombicola using a low-cost
fermentative medium, Appl. Biochem. Biotechnol. 158 (2009) 663–674.
[25] R. Gao, M. Falkeborg, X. Xu, Z. Guo, Production of sophorolipids with
enhanced volumetric productivity by means of high cell density fermentation,
Appl. Microbiol. Biotechnol. 97 (2013) 1103–1111.
[26] Y.B. Kim, H.S. Yun, E.K. Kim, Enhanced sophorolipid production by
feeding-rate-controlled fed-batch culture, Bioresour. Technol. 100 (2009)
6028–6032.
[27] U. Rau, L.A. Nguyen, H. Roeper, H. Koch, S. Lang, Downstream processing of
mannosylerythritol lipids produced by Pseudozyma aphidis, Eur. J. Lipid. Sci.
Technol. 107 (2005) 373–380.
[28] M.C. Cuellar, L.A.M. van der Wielen, Recent advances in the microbial
production and recovery of apolar molecules, Curr. Opin. Biotechnol. 33
(2015) 39–45.
[29] K. Joshi-Navare, P. Khanvilkar, A. Prabhune, Jatropha oil derived
sophorolipids: production and characterization as laundry detergent additive,
Biochem. Res. Int. 2013 (2013) 1–11.

More Related Content

What's hot

Solid dispersions
Solid dispersionsSolid dispersions
Solid dispersionsSwty Sweta
 
Colloidal drug delivery system (Nano formulation)
Colloidal drug delivery system (Nano formulation)Colloidal drug delivery system (Nano formulation)
Colloidal drug delivery system (Nano formulation)pratik9527088941
 
A review on_hydroxyethyl_cellulose
A review on_hydroxyethyl_celluloseA review on_hydroxyethyl_cellulose
A review on_hydroxyethyl_cellulosejiten patel
 
Metoformin HCL - Know how technology
Metoformin HCL - Know how technologyMetoformin HCL - Know how technology
Metoformin HCL - Know how technologyDr.Adel Salem Sultan
 
Solid dispersion technique
Solid dispersion techniqueSolid dispersion technique
Solid dispersion techniqueSidharth Mehta
 
Conference on Co- micronization by PH.D. Jérôme Hecq
Conference on Co- micronization by PH.D. Jérôme HecqConference on Co- micronization by PH.D. Jérôme Hecq
Conference on Co- micronization by PH.D. Jérôme HecqBertin Pharma
 
Classification of Solubility Enhancement Techniques
Classification of Solubility Enhancement TechniquesClassification of Solubility Enhancement Techniques
Classification of Solubility Enhancement TechniquesSantuMistree4
 
SOLID DISPERSION TECHNIQUE
SOLID DISPERSION TECHNIQUESOLID DISPERSION TECHNIQUE
SOLID DISPERSION TECHNIQUERahul Pandit
 
Research Presentation for IESL Annual Sessions 2016
Research Presentation for IESL Annual Sessions 2016Research Presentation for IESL Annual Sessions 2016
Research Presentation for IESL Annual Sessions 2016B.K.T. Samarasiri
 
Methods of enhancing Dissolution and bioavailability of poorly soluble drugs
Methods of enhancing Dissolution and bioavailability of poorly soluble drugsMethods of enhancing Dissolution and bioavailability of poorly soluble drugs
Methods of enhancing Dissolution and bioavailability of poorly soluble drugsRam Kanth
 
Enhancement of dissolution rate and bioavailability of poorly soluble drugs
Enhancement of dissolution rate and bioavailability of poorly soluble drugsEnhancement of dissolution rate and bioavailability of poorly soluble drugs
Enhancement of dissolution rate and bioavailability of poorly soluble drugsNagaraju Ravouru
 
Ijdrt issue march-april-2015_article1
Ijdrt issue march-april-2015_article1Ijdrt issue march-april-2015_article1
Ijdrt issue march-april-2015_article1IJDRT JOURNAL
 
Techniques for enhancement of dissolution rate
Techniques for enhancement of dissolution rateTechniques for enhancement of dissolution rate
Techniques for enhancement of dissolution rateSagar Savale
 
Fabrication and evaluation of a stable flurbiprofen hydrogel
Fabrication and evaluation of a stable flurbiprofen hydrogelFabrication and evaluation of a stable flurbiprofen hydrogel
Fabrication and evaluation of a stable flurbiprofen hydrogelpharmaindexing
 
olanzapine solid dispersion
olanzapine solid dispersion olanzapine solid dispersion
olanzapine solid dispersion Aparna Rajeevi
 
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATION
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATIONSUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATION
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATIONAbu Sayed
 

What's hot (19)

Solid dispersion
Solid dispersionSolid dispersion
Solid dispersion
 
Solid dispersions
Solid dispersionsSolid dispersions
Solid dispersions
 
Colloidal drug delivery system (Nano formulation)
Colloidal drug delivery system (Nano formulation)Colloidal drug delivery system (Nano formulation)
Colloidal drug delivery system (Nano formulation)
 
A review on_hydroxyethyl_cellulose
A review on_hydroxyethyl_celluloseA review on_hydroxyethyl_cellulose
A review on_hydroxyethyl_cellulose
 
Metoformin HCL - Know how technology
Metoformin HCL - Know how technologyMetoformin HCL - Know how technology
Metoformin HCL - Know how technology
 
Solid dispersion technique
Solid dispersion techniqueSolid dispersion technique
Solid dispersion technique
 
Conference on Co- micronization by PH.D. Jérôme Hecq
Conference on Co- micronization by PH.D. Jérôme HecqConference on Co- micronization by PH.D. Jérôme Hecq
Conference on Co- micronization by PH.D. Jérôme Hecq
 
Classification of Solubility Enhancement Techniques
Classification of Solubility Enhancement TechniquesClassification of Solubility Enhancement Techniques
Classification of Solubility Enhancement Techniques
 
SOLID DISPERSION TECHNIQUE
SOLID DISPERSION TECHNIQUESOLID DISPERSION TECHNIQUE
SOLID DISPERSION TECHNIQUE
 
Research Presentation for IESL Annual Sessions 2016
Research Presentation for IESL Annual Sessions 2016Research Presentation for IESL Annual Sessions 2016
Research Presentation for IESL Annual Sessions 2016
 
Methods of enhancing Dissolution and bioavailability of poorly soluble drugs
Methods of enhancing Dissolution and bioavailability of poorly soluble drugsMethods of enhancing Dissolution and bioavailability of poorly soluble drugs
Methods of enhancing Dissolution and bioavailability of poorly soluble drugs
 
Enhancement of dissolution rate and bioavailability of poorly soluble drugs
Enhancement of dissolution rate and bioavailability of poorly soluble drugsEnhancement of dissolution rate and bioavailability of poorly soluble drugs
Enhancement of dissolution rate and bioavailability of poorly soluble drugs
 
Nano suspension copy
Nano suspension   copyNano suspension   copy
Nano suspension copy
 
Ijdrt issue march-april-2015_article1
Ijdrt issue march-april-2015_article1Ijdrt issue march-april-2015_article1
Ijdrt issue march-april-2015_article1
 
CPI 2013
CPI 2013CPI 2013
CPI 2013
 
Techniques for enhancement of dissolution rate
Techniques for enhancement of dissolution rateTechniques for enhancement of dissolution rate
Techniques for enhancement of dissolution rate
 
Fabrication and evaluation of a stable flurbiprofen hydrogel
Fabrication and evaluation of a stable flurbiprofen hydrogelFabrication and evaluation of a stable flurbiprofen hydrogel
Fabrication and evaluation of a stable flurbiprofen hydrogel
 
olanzapine solid dispersion
olanzapine solid dispersion olanzapine solid dispersion
olanzapine solid dispersion
 
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATION
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATIONSUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATION
SUPER CRITICAL FLUID AND THEIR TEXTILE APPLICATION
 

Viewers also liked (18)

Project Presentation Paper
Project Presentation PaperProject Presentation Paper
Project Presentation Paper
 
BALAJI K _Resume
BALAJI K _ResumeBALAJI K _Resume
BALAJI K _Resume
 
ABOUT ME
ABOUT MEABOUT ME
ABOUT ME
 
Sibo CV
Sibo CVSibo CV
Sibo CV
 
Zorgverzekering: hoe zit het?
Zorgverzekering: hoe zit het?Zorgverzekering: hoe zit het?
Zorgverzekering: hoe zit het?
 
HEAA313 Presentation 10230458
HEAA313 Presentation 10230458HEAA313 Presentation 10230458
HEAA313 Presentation 10230458
 
Frances5 france
Frances5 franceFrances5 france
Frances5 france
 
3 profesiones demandadas
3 profesiones demandadas 3 profesiones demandadas
3 profesiones demandadas
 
этнография ансамбль Истоки 2009 Устюг
этнография ансамбль Истоки 2009 Устюгэтнография ансамбль Истоки 2009 Устюг
этнография ансамбль Истоки 2009 Устюг
 
Khumo Seema - social media Fundamentals
Khumo Seema - social media FundamentalsKhumo Seema - social media Fundamentals
Khumo Seema - social media Fundamentals
 
How To Write About Essays
How To Write About EssaysHow To Write About Essays
How To Write About Essays
 
ieeeTemplateWImax1Lab08196366(1)
ieeeTemplateWImax1Lab08196366(1)ieeeTemplateWImax1Lab08196366(1)
ieeeTemplateWImax1Lab08196366(1)
 
Resume
ResumeResume
Resume
 
Tax expenditure on occupational pensions in Ireland: Relevance, Cost & Distri...
Tax expenditure on occupational pensions in Ireland: Relevance, Cost & Distri...Tax expenditure on occupational pensions in Ireland: Relevance, Cost & Distri...
Tax expenditure on occupational pensions in Ireland: Relevance, Cost & Distri...
 
oye oye new
oye oye newoye oye new
oye oye new
 
музейная суббота 2015
музейная  суббота 2015музейная  суббота 2015
музейная суббота 2015
 
Harish Namburi Resume
Harish Namburi ResumeHarish Namburi Resume
Harish Namburi Resume
 
Modelling Increase in Low Pay 18th May 2016
Modelling Increase in Low Pay 18th May 2016Modelling Increase in Low Pay 18th May 2016
Modelling Increase in Low Pay 18th May 2016
 

Similar to integrated sophorolipid production and gravity separation

Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...
Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...
Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...Cátedra Banco Santander
 
Anaerobic methods of waste water treatment v.n.nag
Anaerobic methods of waste water treatment v.n.nagAnaerobic methods of waste water treatment v.n.nag
Anaerobic methods of waste water treatment v.n.nagvishal nag
 
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...IRJET Journal
 
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...IRJET Journal
 
Production of dicarboxylic acid using yeast 1.pptx
Production of dicarboxylic acid using yeast 1.pptxProduction of dicarboxylic acid using yeast 1.pptx
Production of dicarboxylic acid using yeast 1.pptxrajeshkumar590473
 
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...Agriculture Journal IJOEAR
 
Effect of Medium Composition on Changes of Surface Tension During Cultivation...
Effect of Medium Composition on Changes of Surface Tension During Cultivation...Effect of Medium Composition on Changes of Surface Tension During Cultivation...
Effect of Medium Composition on Changes of Surface Tension During Cultivation...IJERA Editor
 
JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023SaraHarmon5
 
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...IJERA Editor
 
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE O
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE OTITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE O
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE OTakishaPeck109
 
bioplastics and biotechnology: Green Plastics.pdf
bioplastics and biotechnology: Green Plastics.pdfbioplastics and biotechnology: Green Plastics.pdf
bioplastics and biotechnology: Green Plastics.pdfRAJESHKUMAR428748
 
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...ijtsrd
 
Paper2. Reanto
Paper2. ReantoPaper2. Reanto
Paper2. ReantoZeban Shah
 
Extraction of Bio-Fuel from Algae by Anaerobic Digestion
Extraction of Bio-Fuel from Algae by Anaerobic DigestionExtraction of Bio-Fuel from Algae by Anaerobic Digestion
Extraction of Bio-Fuel from Algae by Anaerobic DigestionEditor IJMTER
 

Similar to integrated sophorolipid production and gravity separation (20)

Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...
Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...
Promoviendo una educación multicultural e interdisciplinar: Químicos Británic...
 
Anaerobic methods of waste water treatment v.n.nag
Anaerobic methods of waste water treatment v.n.nagAnaerobic methods of waste water treatment v.n.nag
Anaerobic methods of waste water treatment v.n.nag
 
Formulationassessment
FormulationassessmentFormulationassessment
Formulationassessment
 
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...
IRJET-Influence of Advanced Settling Zone on COD Removal Efficiency of UASB R...
 
SOIL BURIAL DEGRADATION OF POLYPROPYLENE/ STARCH BLEND
SOIL BURIAL DEGRADATION OF POLYPROPYLENE/ STARCH BLENDSOIL BURIAL DEGRADATION OF POLYPROPYLENE/ STARCH BLEND
SOIL BURIAL DEGRADATION OF POLYPROPYLENE/ STARCH BLEND
 
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...
Biosynthesis of Polyhydroxyalkanoate blends from Cassia Seed: GalactoMannan a...
 
Production of dicarboxylic acid using yeast 1.pptx
Production of dicarboxylic acid using yeast 1.pptxProduction of dicarboxylic acid using yeast 1.pptx
Production of dicarboxylic acid using yeast 1.pptx
 
OPB__Prashant Patel
OPB__Prashant PatelOPB__Prashant Patel
OPB__Prashant Patel
 
PMG Newsletter (Volume 03. Issue 16).pdf
PMG Newsletter (Volume 03. Issue 16).pdfPMG Newsletter (Volume 03. Issue 16).pdf
PMG Newsletter (Volume 03. Issue 16).pdf
 
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
 
Bio-substitution
Bio-substitutionBio-substitution
Bio-substitution
 
Effect of Medium Composition on Changes of Surface Tension During Cultivation...
Effect of Medium Composition on Changes of Surface Tension During Cultivation...Effect of Medium Composition on Changes of Surface Tension During Cultivation...
Effect of Medium Composition on Changes of Surface Tension During Cultivation...
 
JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023
 
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...
Mixotrophic Cultivation of Botryococcus Braunii for Biomass and Lipid Yields ...
 
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE O
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE OTITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE O
TITLE PAGETABLE OF CONTENTSContentsTITLE PAGE1TABLE O
 
bioplastics and biotechnology: Green Plastics.pdf
bioplastics and biotechnology: Green Plastics.pdfbioplastics and biotechnology: Green Plastics.pdf
bioplastics and biotechnology: Green Plastics.pdf
 
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...
Effect of Glucose on Biosurfactant Production using Bacterial Isolates from O...
 
Ae35171177
Ae35171177Ae35171177
Ae35171177
 
Paper2. Reanto
Paper2. ReantoPaper2. Reanto
Paper2. Reanto
 
Extraction of Bio-Fuel from Algae by Anaerobic Digestion
Extraction of Bio-Fuel from Algae by Anaerobic DigestionExtraction of Bio-Fuel from Algae by Anaerobic Digestion
Extraction of Bio-Fuel from Algae by Anaerobic Digestion
 

integrated sophorolipid production and gravity separation

  • 1. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 Process Biochemistry xxx (2016) xxx–xxx Contents lists available at ScienceDirect Process Biochemistry journal homepage: www.elsevier.com/locate/procbio Integrated sophorolipid production and gravity separation Ben M. Dolman, Candice Kaisermann, Peter J. Martin, James B. Winterburn∗ School of Chemical Engineering and Analytical Science, The Mill, The University of Manchester, Manchester, M13 9PL, UK a r t i c l e i n f o Article history: Received 9 November 2016 Received in revised form 13 December 2016 Accepted 21 December 2016 Available online xxx Chemical compounds: Sophorolipid (PubChem CID: 11856871) Keywords: Integrated separation Sophorolipid Settling Candida bombicola Biosurfactant Glycolipid a b s t r a c t A novel method for the integrated gravity separation of sophorolipid from a fermentation broth has been developed, enabling removal of a sophorolipid phase of either higher or lower density than the bulk fer- mentation broth, while cells and other media components are recirculated and returned to the bioreactor. The capability of the separation system to recover an enriched sophorolipid product phase was demon- strated on three sophorolipid producing fed batch fermentations using Candida bombicola, giving an 11% reduction in fermenter volume required whilst maintaining sophorolipid production. Sophorolipid recov- eries of up to 86% (280 g) of the total produced over a whole fermentation were achieved at an enrichment of up to 9. Furthermore, the broth viscosity reduction achieved by removal of the sophorolipid phase enabled a 34% reduction in mixing power to maintain the same dissolved oxygen level by the end of the fermentation, with a 9% average reduction over the course of the fermentation. Fermentation duration could be extended to 1023 h, allowing production of 623 g sophorolipid from 1 l initial batch volume. These benefits could lead to a substantial decrease in the cost of sophorolipid production, making high volume applications such as enhanced oil recovery economically feasible. © 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 1. Introduction Sophorolipids are microbially produced glycolipid biosurfac- tants, which are rapidly increasing their market share of the 27 billion USD global surfactant market [1]. While several yeast strains are able to synthesize sophorolipids, most research and industrial use is focused on Candida bombicola ATCC 22214, the organism used in this study [2]. Sophorolipids consist of a hydrophilic sophorose disaccharide bound to a hydrophobic fatty acid with a typical chain length of 16–18 carbon atoms. The fatty acid may be joined by an ester bond to the second glucose monomer, giving a lactonic sophorolipid, or joined to only one glucose monomer, giving an acidic sophorolipid due to the unbound fatty acid. These and other differences in the fatty acid chain and acetylation of the sophorose molecules give a range of different structures and properties. Two common structures representing lactonic and acidic sophorolipids are shown in Fig. 1 [2]. Sophorolipids are produced industrially by a number of compa- nies, who often utilize sophorolipids’ detergent and low foaming properties in a variety of formulated cleaning products [3]. The therapeutic properties of sophorolipids have allowed them to be commercialized in anti-dermatitis soap and other body washes, ∗ Corresponding author. E-mail address: james.winterburn@manchester.ac.uk (J.B. Winterburn). and in a cream to reduce oily skin by MG Intobio Co and Soliance. There is ongoing research into potential medical applications of sophorolipid, with anti-cancer, anti-HIV, antimicrobial and anti- biofilm activity being investigated [4–6]. Sophorolipids also have potential for use in low cost, high volume applications such as bioremediation and enhanced oil recovery if production costs can be significantly reduced [7,8]. In sophorolipid producing fermentations, product concentra- tions of over 300 g l−1, with productivities of over 1 g l−1 h−1 are routinely achieved [9–11]. Sophorolipid producing fermentations begin with a cell growth phase, which typically lasts until the nitrogen in the media is depleted, at which point the sophorolipid production rate increases significantly, if both a hydrophilic and a hydrophobic carbon source are present [12]. The sophorolipid pro- duction phase normally lasts for around 200 h, at which point the dissolved oxygen level in the fermenter cannot be maintained due to oxygen mass transfer limitation. This dissolved oxygen reduction is caused by the high viscosity of the sophorolipid produced, meaning the fermentation must be stopped and the sophorolipid recovered [12,13]. It is well known that the presence of a separate sophorolipid phase in the bioreactor significantly reduces the oxygen mass transfer coefficient, kLa, by both providing a resistance to mass transfer across the air/liquid interface and increasing the viscosity of the medium, which results in oxygen limitation, increased stirring power requirements and non-homogeneity in the bioreactor [12–14]. http://dx.doi.org/10.1016/j.procbio.2016.12.021 1359-5113/© 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
  • 2. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 2 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx Fig. 1. Molecular structure of common lactonic (left) and acidic (right) sophorolipids. A lactonic bond can be seen joining the fatty acid chain to the second glucose monomer of the sophorose in the lactonic sophorolipid, where the acidic sophorolipid has a free carboxylic acid to end its fatty acid chain. The physical form of sophorolipids is dependent on the con- ditions under which they are produced, which directly affect the proportions of acidic and lactonic sophorolipids produced. Sophorolipids typically separate from the fermentation broth as a crystalline material if the lactonic to acidic ratio is high and the hydrophobic carbon source concentration is low. The sophorolipids otherwise form a viscous second phase of around 50% sophorolipid and 50% water, which may sit below residual oil at the surface of the broth or sink to the bottom of the bioreactor when agitation is stopped [10,15,16]. These properties are commonly exploited at the end of a fermen- tation to give an easy, crude separation of the sophorolipid from the fermentation broth, either by crystal decantation, crystal filtration or decantation of the sophorolipid gel [11,16,17]. These techniques have not previously been used effectively to recover sophorolipids during fermentation. Industrially, there are a number of costs associated with repeated batch cycles for sophorolipid fermentation, in terms of downtime between cycles, cleaning costs and the lengthy inocu- lum preparation required for large scale production [18,19]. There are numerous proposed partial solutions to this problem in the lit- erature. For example, a portion of the broth can be removed and replaced with fresh media, allowing high productivity to be main- tained for seven 80–130 h cycles, nevertheless removing biomass proportionally to other components [19]. Sophorolipid settling by gravity within a fermentation vessel or shake flask has previously been demonstrated for small scale sophorolipid production. Signif- icant benefits of sophorolipid separation have been shown, with a doubling of the duration of sophorolipid production and little effect on production after 15 min without agitation or aeration demon- strated by Guilmanov et al. [20], and a productivity increase from 1.38 to 1.89 g l−1 h−1 shown by Marchal et al. [19]. Both studies rely on gravity settling within the fermentation vessel or shake flask, however, making scale up impractical due to the excessive settling distances present if this technique were applied at industrial scale. Effective integrated separation techniques have been developed for other biosurfactant systems, notably foam fractionation for hydrophobin proteins, surfactin and rhamnolipids, but there have been no successful scalable attempts at integrated separation for sophorolipid production [21–23]. This paper details a novel technique, based on an integrated gravity settling column, for removing the sophorolipid phase from the fermentation broth during fermentation, reducing the fermen- tation volume required which allows continued substrate feeding and provides a concentrated product phase. We also demonstrate, for the first time, the application of this technique to extend the production phase of a fermentation beyond 1000 h, significantly increasing batch production. 2. Methodology 2.1. Fermentation Sophorolipid was produced by fed batch fermentation using C. bombicola ATCC 22214 in an Electrolab Fermac 320 fermentation system (Electrolab, UK) with a 2 l maximum working volume, H:D of 2, two 55 mm diameter 6-bladed Rushton-type impellers and an initial working volume of 1 l, with a stirring rate of 200–800 rpm. This was sufficient to disperse vegetable oil within the bioreactor. Growth medium for the fermentations, preculture and agar plates contained 6 g l−1 yeast extract and 5 g l−1 peptone. The initial con- centration of glucose in all fermentations preculture and agar plates was 100 g l−1, with an initial rapeseed oil concentration of 50 g l−1 in the fermenters, 100 g l−1 in the preculture and 0 g l−1 in the agar plates. Four fermentations were carried out. Fermentation 1 was con- ducted in a conventional manner without separation. Fermentation 2 was directly comparable to fermentation 1 except that the in situ separator was used to remove product from the top of the separa- tor. Fermentation 3 was controlled to give a product that separated from the bottom of the separator. Fermentation 4 was carried out for an extended period of time to demonstrate the potential to utilise the integrated separation to extend the fermentation period and give higher batch production, and controlled to give separation from the surface of the broth. C. bombicola was first transferred from cryogenic storage (−80 ◦C) onto agar plates, and incubated at 25 ◦C for 48 h. Single colonies from these plates were then used to inocu- late 50 ml of medium in 250 ml shake flasks, which were incubated at 25 ◦C and 200 rpm for 30 h. This inoculum was diluted to an opti- cal density of 20 at 600 nm with fresh media and 100 ml used to inoculate the fermenter. Fermentations were run at 25 ◦C, and dissolved oxygen was con- trolled to 30% by varying the stirrer speed, whilst maintaining a
  • 3. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 3 constant aeration rate of 1 l min−1. Fermenter pH was controlled to a value of 3.5 by the addition of 3 M sodium hydroxide. The feeding rates of rapeseed oil and glucose were similar for fermentation one and two, to facilitate comparison, with different feeding rates used for fermentation three and four, to give separa- tion of sophorolipid to the top and bottom of the fermenter. Feeding rates of oil were modified during the experiments to maintain a low concentration, without limiting production, according to the oil concentration in the samples. Glucose concentration was used to control the relative density of the sophorolipid phase and the bulk media to enable effective separation from either the top or bottom of the separator, as well as being an important substrate for sophorolipid production. The feed profiles are shown in Fig. 4. The total sophorolipid produced was calculated by adding the mass of sophorolipid in the fermenter and the mass of sophorolipid removed from the fermenter using the separator. Contamination was tested for visually using microscopy and by streak plating. 2.2. Separation Integrated separation of the sophorolipid from the fermentation broth was carried out using an in house built settling column, as shown in Fig. 2. The integrated arrangement of the bioreactor and settling column is shown in Fig. 3 with the settling column sup- ported at an angle of 30◦ from horizontal. The separator was rinsed with 70% ethanol before attaching to the fermenter. Sophorolipid separation was carried out intermittently, based on visual observa- tion of a sample taken from the bioreactor. When a significant layer of sophorolipid rich phase could be seen at either the top or bot- tom of the sample in the universal bottle within 2 min separation was initiated. Prior experiments indicated that at these conditions separation is effective. During separation, which was used intermittently during fer- mentation, sophorolipid rich fermentation broth was continuously circulated from the fermenter, through the settling column and back to the fermenter. This was pumped from the fermenter through a stainless steel tube with an inlet 20 mm from the bot- tom of the fermenter, and then in 8 mm external diameter silicon tubing of 1 mm wall thickness to and from the separator. The flow rate of media into and out of the settler was controlled to 1 ml s−1 using Matson Marlow 502 S and 503 U pumps (Watson Marlow, UK). This flowrate was based on the results of preliminary experi- ments, giving a residence time in the settling column of 76 s, with a total residence time in the column and tubing of 137 s. In the set- tling column, the sophorolipid phase separates out towards either the top or bottom of the column, depending on the relative density of the sophorolipid and bulk media. Initially, broth is continuously circulated and the sophorolipid product collects in the settling col- umn. When the sophorolipid phase accumulating in the separator reached 50% of the height of the separator, which typically occurred after around three minutes of separator operation, the outlet pump was started to continuously remove the sophorolipid product phase at a rate controlled between 0.5 and 2 ml min−1, depending on the accumulation or reduction of the sophorolipid phase in the set- tling vessel. The separation was stopped when the separation rate dropped below 0.5 ml min−1, until the condition for separation was again observed. 2.3. Hydrodynamics To determine the effect of the sophorolipid phase on the agita- tion requirements, the sophorolipid enriched fractions, which had been separated from the bioreactor using the separator over the course of fermentation 2, were pooled and returned to the fer- menter at 308 h after the start of the fermentation over a period of 12 min. The stirrer speed and dissolved oxygen percentages were then monitored as the fermentation control returned the dissolved oxygen percentage to the set point. The equation for power number is shown in Eq. (1); P = Np n3 D5 (1) where P is power, Np is power number, is density, n is stirrer speed and D is stirrer diameter. Due to identical fermenters being used and assuming the den- sity is constant (as density changes during the fermentation are relatively small), the power input as a function of power number can be calculated during the fermentation. The power input was integrated over time to determine the power input for the whole fermentation, with the power input for fermentations equated until the time point of the first separation. Separation performance was measured in terms of enrichment and recovery, which are defined in Eqs. (2) and (3); enrichment = Cp Cf (2) recovery (%) = Cp × Vp Cf × Vf × 100 (3) where Cp is the sophorolipid concentration in the product, Cf is the product concentration in the fermenter before separation, Vp is the volume of product phase recovered, and Vf is the initial volume of liquid in the fermenter. 2.4. Analytical techniques For all analysis, 5 ml of broth was removed from the bioreactor or 5 ml product was taken from the sophorolipid collection vessel con- nected to the separator. The sample was centrifuged at 5000 rpm for 5 min using a Sigma 6–16S centrifuge (Sigma laboratory cen- trifuges, Germany) and the glucose in the supernatant quantified using a TrueResult ® blood glucose monitor (Nipro, Japan). A hexane extraction to extract residual rapeseed oil fol- lowed by a triple ethyl acetate extraction to extract sophorolipid were then applied to the whole sample, with oil concentra- tion measured gravimetrically from the hexane extraction, and sophorolipid measured gravimetrically from the pooled ethyl acetate extracts[24–26]. These extracts were dried to constant weight in weighing dishes at ambient temperature for 30 h. Cell growth was determined by both dry cell weight and optical density measurement. After the aforementioned hexane and ethyl acetate extractions, 8 ml distilled water was added to the remainder of the sample in the centrifuge tubes, which were then centrifuged at 8000 rpm for 10 min. The supernatant was discarded and the resulting cell pellet was resuspended in 8 ml distilled water. This cell suspension was transferred to drying trays, which were dried to constant weight at 90 ◦C in a drying oven. Optical density was used as a proxy for dry cell weight when diluting the inoculum, at a wavelength of 600 nm. The structure of the sophorolipids produced was determined with negative electrospray ionisation mass spectrometry, using an Agilent 6520 QTOF mass spectrometer (Agilent, United States). Samples were prepared by redissolving the ethyl acetate extracts, i.e. sophorolipids, in ethyl acetate, and filtering using a 0.2 ␮m filter. Flow injection analysis was used, at 0.3 ml min−1, 50% acetonitrile, 0.1% formic acid and 49.9 % water, with an injection volume of 2 ␮l. The viscosity of the product phase was measured using an AR2000 controlled rotational rheometer with cone geometry (TA Instruments, USA).
  • 4. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 4 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx Fig. 2. Diagram of custom built sophorolipid separator used for this study. Plan view, side view and end view are shown, all dimensions are internal dimensions shown in mm.. 3. Results and discussion A novel gravity sophorolipid separation technique was success- fully applied to three fermentations. Results from one fermentation without separation, fermentation 1, are presented alongside results from three fermentations with separation, fermentations 2, 3 and 4. Fermentations 1 and 2, without and with integrated separa- tion respectively, are directly comparable, to allow evaluation of the effect of separation, but due to feeding rate control to enable sophorolipid to be recovered from the bottom of the separator fermentation 3 is not directly comparable. Fermentation 4 was intended to demonstrate an extended fermentation, and so is also not directly comparable. The separation was run periodically in all fermentations, when sufficient sophorolipid phase had accu- mulated, with the majority of the available sophorolipid phase separated. 3.1. Fermentations Fig. 4 shows the feeding rate of substrates for the fermentations presented, which enabled control of the sophorolipid phase to sep- arate from the surface or bottom of the separator, whilst also being an important parameter for sophorolipid production. The progress of the fermentations over time is presented in Fig. 5, and the key metrics from these fermentations are presented in Table 1. Fig. 5 shows the progress of fermentation 1, without separation, and fermentations 2, 3 and 4, during which sophorolipid product was separated from the fermentation broth. In fermentation 2 and 4, sophorolipid was recovered at the top of the integrated gravity separator, and in fermentation 3 the sophorolipid was collected from the bottom of the separator. In fermentation 2, separation was carried out at 111, 184 and 261 h, in fermentation 3 at 72, 281, 355 and 376 h and in fermentation 4 at 86, 111, 160, 186, 232 and 540 h. No separation of the sophorolipid phase was carried out in fermentation 1. In fermentation 2, the glucose concentration initially rose, and remained above 50 g l−1 for the majority of the fermentation, which led to the sophorolipid rising to the surface of the fermentation broth without agitation. A high glucose concentration throughout Table 1 Key metrics for all sophorolipid producing fermentations. Fermentation 2, 3 and 4 used sophorolipid separation, with fermentation 1 included for comparison. All metrics are the result of unique fermentations. Due to volume changes caused by substrate addition and product removal, productivity is based on the initial fer- menter working volume and the total sophorolipid produced. Fermentation 1 2 3 4 Product separation none top bottom top Duration (h) 305 305 379 1023 Yield substrate consumed (g g−1 ) 0.43 0.53 0.42 0.53 Yield substrate fed (g g−1 ) 0.33 0.37 0.39 0.47 Productivity (g l−1 h−1 ) 1.07 1.07 0.71 0.61 Maximum fermenter volume (l) 1720 1544 1350 1550 Total sophorolipid produced (g) 325 325 270 623 Total dry cell weight (g) 16.1 21.0 16.5 32.1 Yield product on biomass (g g−1 ) 20.2 15.5 16.4 19.4 fermentation 4 meant the sophorolipid was also separated from the top of the separator during this experiment. Sophorolipid was first separated from the bottom of the sepa- rator at 71.5 h in fermentation 3, when a sophorolipid phase could be observed to settle in a sample bottle within 2 min. Settling was not possible after this until 283 h due to the high residual glucose concentrations caused by pulse glucose feeding, which was used to ensure good sophorolipid production. Whilst the relative density of the sophorolipid phase and the broth also depend on other factors, a glucose concentration of 50 g l−1 tends to represent a threshold of sophorolipid phase separation to the surface or the bottom of the separator. This is because higher glucose concentrations lead to higher media densities, meaning the sophorolipid phase is rela- tively less dense the higher the glucose concentration. After 283 h the glucose concentration had dropped sufficiently for settling to be used again. Lower glucose feeding could enable sophorolipid settling throughout the fermentation, though this might impact sophorolipid production. Fermentations 1 and 2, which were identical apart from the application of integrated sophorolipid separation in fermentation 2, each produced 325 g sophorolipid, with 270 g sophorolipid pro- duced during fermentation 3. The dry cell weight production, of 16–21 g, and the yields of product on substrate, of 0.33-0.39 are
  • 5. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 5 Fig. 3. Integrated fermentation system. Bioreactor is shown on the left, with the separator in the center, and the product collection vessel on the right. Broth is pumped from the bioreactor into the separator, and recirculated back to the bioreactor. Product is pumped from the separator into the collection vessel. This can be used for; (a) − sophorolipid phase density higher than fermentation broth. (b)− sophorolipid phase density lower than fermentation broth. in line with other results in the literature, as are the values for total sophorolipid produced [9,25]. The use of the separator appears to have little effect on the total production of sophorolipids over the same time period, with identical values recorded for the two comparable fermentations. In fermentation 2 with separation, 21 g of biomass were pro- duced, more than the 16 g of cell biomass produced in fermentation 1, with product yields on biomass of 15.5 g g−1 and 20.2 g g−1 respectively. The reason for the reduction in product yield on biomass in fermentations 2 and 3 is not known, but with further optimisation fermentation 4, with separation, reached a product yield on biomass of 19.4 g g−1. The highest working volume reached during a fermentation dic- tates the overall fermenter volume required, and the high feeding rates used during sophorolipid producing fermentations lead to a large increase in the working volume required over time, much of which is only required in the later stages of the fermentation. The use of integrated sophorolipid separation in fermentation 2 decreased the fermenter working volume required by removing 523 ml broth from the fermenter using the separator. This meant only 1540 ml working volume was needed for fermentation 2 com- pared to 1720 ml in fermentation 1 without separation. This 11% decrease in volume requirement could reduce bioreactor capital costs. The corresponding maximum volume for fermentation 3 was 1350 ml, but was not directly comparable due to differences in feeding rates between fermentations 2 and 3. The overall productivity of fermentations 1 and 2 were 1.07 g l−1 h−1 calculated at the starting volume, with a correspond- ing productivity of 0.77 g l−1 h−1 for fermentation 3. The rate of sophorolipid production slowed dramatically when the oil was depleted after around 80 h in fermentations 1–3, reducing from 2 g l−1 h−1 to 0.6 g l−1 h−1 and increasing the oil feed rate did not return the productivity to the previous level. Many studies have demon- strated a fairly constant production rate throughout a fermentation until the point at which the fermentation had to be stopped due to dissolved oxygen limitation, and with improved feeding control in fermentation 4, a productivity of 2.0 g l−1 h−1 was maintained until 158 h [17] [12]. In fermentation 4, fermentation was continued past the point at which fermentations usually have to be stopped due to product
  • 6. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 6 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx Fig. 4. Feeding profiles of glucose and rapeseed oil for all fermentations in this study. Glucose (blue), rapeseed oil (green) and total (red) shown for fermentation 1 (solid, a) fermentation 2 (dotted, a), fermentation 3 (b) and fermentation 4 (c). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) accumulation, and run for a total of 1023 h. This enabled the produc- tion of 623 g sophorolipid from a 1 l initial broth, which compares favourably to the highest previous reported titers of around 400 g l−1, and clearly demonstrates the capacity of integrated separation to extend the time period of sophorolipid production in fermenta- tion. This could lead to a dramatic improvement in overall process productivity, by reducing the proportion of time spent in inoculum preparation, biomass production and cleaning. 3.2. Separation The separation results achieved in fermentations 2, 3 and 4 are shown in Table 2. During fermentations 2, 3 and 4 with integrated separation, the majority of the sophorolipid was removed from the fermentation broth, with 86% of the total sophorolipids produced separated in fermentation 2, 74% separated in fermentation 3 and 65% separated in fermentation 4. Almost no cells and only 8 g of oil were removed by the separa- tion over the course of fermentation 3, determined by gravimetric analysis as for fermentation samples. This is because the rate of settling of the cells was much slower than the settling of the sophorolipid product, and oil rose to the surface of the separa- tor rather than sinking to the bottom of the separator with the sophorolipid. Cell removal was also negligible in fermentation 2, though 68 g of oil was removed, which was reduced by better feed- ing rate control in fermentation 4, where only 20 g oil was removed, while again separating from the surface of the separator. 2.6 g cells were removed during fermentation 4, which represents a small proportion of the total biomass, 32.1 g. The enrichment varied significantly between separations at different time points, from 2.5 to 9. This is largely due to the sophorolipid concentration present in the fermenter before the separation, as there was little variation of the concentration in the sophorolipid enriched product fraction, of approximately 550 g l−1. In fermentation 2, when the sophorolipid phase was separated from the top of the settler, a total of 68 g oil was recovered along with the sophorolipid during the fermentation, which was the primary reason for the variations in the concentration of the sophorolipid phase. There is little scope to improve the product phase concentra- tion above that demonstrated in fermentation 3 using the current technique, however, with an increased fermenter volume, the sys- tem could operate at lower initial sophorolipid concentrations and so give an improved enrichment. Sophorolipid recoveries would be expected to improve sig- nificantly as fermentation scale is increased; in laboratory scale experiments the separation had to be stopped as the layer of sophorolipids at the bottom/top of the settling column became too low, to prevent the media and cell phase being entrained in the product stream. This minimum sophorolipid phase depth, which
  • 7. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 7 Fig. 5. Time course of fermentations using sophorolipid separation. Results for fermentation 1 (a), fermentation 2 (b), fermentation 3 (c) and fermentation 4 (d) are presented. Dry cell (black squares) glucose (blue triangles) rapeseed oil (green circles) and sophorolipid (filled red circles) concentrations are shown. Arrows show sophorolipid separation, total sophorolipid produced shown by open red circles. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) Table 2 Separation results for fermentation 2, fermentation 3 and fermentation 4. All metrics are the result of unique fermentations, with the application of the novel integrated gravity separation developed in this study. Time (h) Sophorolipid recovered (g) Sophorolipid concentration (g l−1 ) Total sophorolipid present (g) Sophorolipid product concentration (g l−1 ) Enrichment Recovery at time point (%) Fermentation 2 111 97.1 147.7 259.5 550.6 3.73 37 184 99.2 118.0 165.2 461.6 3.91 60 261 83.8 89.2 96.4 540.9 6.07 87 Total 280.1 86 Total oil removed by separation (g) 8 Total cells removed by separation (g) below detection limit Fermentation 3 71.5 16.8 103.7 128.4 582.9 5.62 13 281 79.5 168.3 238.2 654.1 3.89 33 355 59.2 109.2 175.3 616.9 5.66 34 376 45.5 106.3 148.9 638.7 6.01 31 Total 201.0 74 Total oil removed by separation (g) 68 Total cells removed by separation (g) below detection limit Fermentation 4 85.6 93.1 152.7 229.1 443.5 2.90 40 111.3 53.5 97.1 127.8 504.8 5.20 42 159.7 61.9 129.6 186.8 538.1 4.15 33 185.8 57.9 72.5 105.3 567.8 7.83 55 231.5 48.9 51.7 68.3 465.7 9.01 72 540.3 89.0 124.7 191.3 312.2 2.5 47 total 404.3 65 Total oil removed by separation (g) 20 Total cells removed by separation (g) 2.6 must be recycled back to the fermenter, would be identical irre- spective of fermenter volume while maintaining a given size of separator, hence a larger total fermenter volume would lead to a large reduction in residual sophorolipid concentration.
  • 8. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 8 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx Fig. 6. The effect of sophorolipid separation on agitation requirements. Dissolved oxygen and stirrer speed profiles showing the increase in stirrer speed required to maintain the dissolved oxygen concentration after sophorolipid rich fractions separated during fermentation 2 returned to fermenter in fermentation 2 at 308 h. Whilst the separator was designed for continuous operation hydrodynamic considerations, in particular the turbulence caused by inlet and outlet disturbances which become more significant with decreasing scale, mean it could not be scaled down further to match the sophorolipid production rates achieved in the 1 l initial working volume fermenter. At the scale presented in this paper, the separation occurred at a rate of around 2 ml min−1, or around 1 g min−1 sophorolipid, 30–150 times greater than the production rate. This separation rate makes it suitable for use with a separa- tor of 30–60 l volume for continuous separation of the sophorolipid produced. Product recovery rate is expected to scale proportionally to the volume of the separating column, so for a new settler design volume can be increased while maintaining or reducing the ratio of inertial to viscous forces, i.e. the Reynolds number. Residence time should be increased proportionally to diameter increase, to allow the same quantity of sophorolipid to settle or float to the surface at the same settling/rising rate. The system could easily be connected for steam in place sterilization at industrial scale. This is the first study to present the design and demonstrate the feasibility of a separation system for sophorolipid production that could be continuously applied, having been shown in this manuscript to separate sophorolipid whilst production continues in the bioreactor for periods of more than one hour. It is also the first integrated separation system applicable to large scale fermen- tation, because it does not rely on separation within the bioreactor, and therefore the first to enable an extended sophorolipid produc- tion period at scale. The reduced bioreactor volume, and reduced start up and cleaning costs of this system could significantly improve the economics of sophorolipid production by increasing the total sophorolipid produced per batch. This would make bulk application, such as for enhanced oil recovery, a more realistic proposition. Other glycolipid biosurfactants, notably rhamnolipids and man- nosylerythritol lipids (MELs), as well as many other bioproducts including those used as biofuels, may also form a separate, insol- uble, phase in a fermentation broth, and so the gravity separation technique presented in this paper could likely also be applied to these systems [27,28]. 3.3. Effect on hydrodynamics and mass transfer Fig. 6 shows the dissolved oxygen level and stirrer speed at the end of fermentation 2, capturing the addition of the sophorolipid rich fractions which were removed by separation during fermen- tation 2 and subsequently pooled, and added to the bioreactor at 308 h. The presence of this sophorolipid phase effectively reduced the volumetric oxygen transfer coefficient, kLa, in the fermenter, due to its high viscosity of around 0.5 Pa.s, resulting in an increase in stirrer speed to maintain the dissolved oxygen at the set point. A stirring rate increase of around 75 rpm, from 500 rpm to 575 rpm was required to maintain the desired dissolved oxygen level when the sophorolipid product separated over the whole fermentation Fig. 7. Mass spectrum of sophorolipid from the end of fermentation 2. Key peaks of C18:1 diacylated acidic sophorolipid at m/z 705 and C18:1 lactonic sophorolipid at m/z 687.
  • 9. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx 9 was added. Given the high viscosity of the sophorolipid phase, it is possible that turbulence was not attained even at the higher agitation rate; in this instance a still higher agitation rate would be required for proper dissolved oxygen control throughout the bioreactor, leading to larger savings than calculated. The mixing Power number in the turbulent regime is typically constant, so changes in impeller speed have a cubic impact on the mixing power. Removing the sophorolipid phase resulted in a 13% decrease in impeller speed and a 34% decrease in mixing power requirement by the end of the fermentation. The relative power requirements over fermentation 1 and 2 were also compared, with an 18% reduction in power input from the time point of the first separation observed, and a 9% improvement when the whole fermentation is taken into account. There would be some increased power consumption from the pumping of the fermentation broth between the separator and the fermenter, but it is expected this would be significantly smaller than the agitation power at scale, as relatively low pumping flow rates are used. If these power reductions can be achieved at indus- trial scale, they can give significant cost savings. 3.4. Sophorolipid structure Mass spectra of sophorolipid samples were taken to determine the sophorolipid structures produced during the fermentation, with the mass spectrum of the sophorolipids taken from the end of fermentation 2 shown in Fig. 7. The main peak at m/z=705 represents a diacylated acidic C18:1 sophorolipid, with the peak at 687 representing a diacylated C18:1 lactonic sophorolipid [29]. There were some differences in the ratios of peak heights between sophorolipid taken from the settler and sophorolipid taken from the fermentation broth immediately before, suggesting this tech- nology may have potential to act as a crude separator of different sophorolipid forms. Recent research has revealed the enzyme responsible for lac- tonisation of acidic sophorolipids, enabling the use of genetic engineering of the genome of Starmerella bombicola, and robust production and purification techniques to yield a 98% pure lactonic or acidic sophorolipid, making the application of sophorolipid in medicinal or personal care products much more likely [3]. Comple- mentary to the advances in genetic engineering and purification, this work makes significant progress in solving some of the complex engineering challenges involved with sophorolipid production, giv- ing potentially dramatic reductions in production cost and making large scale application of sophorolipid more feasible. 4. Conclusions A novel method for the integrated separation of sophorolipid from a fermentation process has been developed. The design of the system overcomes the production and processing difficulties asso- ciated with in situ (i.e. in the fermenter vessel) gravity separation of sophorolipids for scale up, and with a separator residence time of less than two minutes the process seemed to have no impact on further sophorolipid production by the cells. A sophorolipid phase can be removed from the fermentation broth if the product phase had higher or lower density than the media, with enrichments of up to 9, an overall recovery of 86%, and up to 404 g of sophorolipid recovered from the fermentation broth. We demonstrated an 11% decrease in bioreactor working vol- ume requirement when using the separator, due to the removal of the sophorolipid product phase. Optimised feeding rates and settler usage could further reduce the volume requirement. By using the separator, the fermentation could be run for 1023 h, and produce 623 g sophorolipid from a 1 l initial batch, reducing the number of fermentations required for a given product mass. An 18% average reduction in stirrer power was demonstrated over the course of a fermentation once sophorolipid separation was initiated, which translates to around 9% over the entire fermenta- tion, with a 34% decrease in power input shown by the end of the fermentation. The integrated separation system presented in this paper has been developed for sophorolipid separation, but could equally be applied to the production of other insoluble bioproducts, in par- ticular mannosylerythritol lipids. With correct scaling up, it is anticipated that the advantages this system offers will lead to a dramatic improvement of the economics of sophorolipid produc- tion. Acknowledgements Sara Bages Estopa is acknowledged for her invaluable comments on the paper, and Reynard Spiess and Shaun Leivers are acknowl- edged for their help with mass spectrometry. The technology described in this manuscript has been filed for a patent entitled ‘Improvements in and related to lipid production’. The authors are grateful for financial support from the UK Engi- neering and Physical Sciences Research Council (EP/I024905) and the EPSRC DTA fund, which enabled this work to be conducted. References [1] R. Geys, W. Soetaert, I. Van Bogaert, Biotechnological opportunities in biosurfactant production, Curr. Opin. Biotechnol. 30 (2014) 66–72. [2] I. Van Bogaert, K. Saerens, C. De Muynck, D. Develter, W. Soetaert, E.J. Vandamme, Microbial production and application of sophorolipids, Appl. Microbiol. Biotechnol. 76 (2007) 23–34. [3] S.L.K.W. Roelants, K. Ciesielska, S.L. De Maeseneire, H. Moens, B. Everaert, S. Verweire, Q. Denon, B. Vanlerberghe, I. Van Bogaert, P. Van der Meeren, B. Devreese, W. Soetaert, Towards the industrialization of new biosurfactants: biotechnological opportunities for the lactone esterase gene from Starmerella bombicola, Biotechnol. Bioeng. 113 (2016) 550–559. [4] M.A. Díaz De Rienzo, I.M. Banat, B. Dolman, J. Winterburn, P.J. Martin, Sophorolipid biosurfactants: possible uses as antibacterial and antibiofilm agent, N. Biotechnol. 32 (2015) 720–726. [5] V. Shah, G.F. Doncel, T. Seyoum, K.M. Eaton, I. Zalenskaya, R. Hagver, A. Azim, R. Gross, Sophorolipids, microbial glycolipids with anti-human immunodeficiency virus and sperm-immobilizing activities, Antimicrob. Agents Chemother. 49 (2005) 4093–4100. [6] L. Shao, X. Song, X. Ma, H. Li, Y. Qu, Bioactivities of sophorolipid with different structures against human esophageal cancer cells, J. Surg. Res. 173 (2012) 286–291. [7] A.E. Elshafie, S.J. Joshi, Y.M. Al-Wahaibi, A.S. Al-Bemani, S.N. Al-Bahry, D. Al-Maqbali, I.M. Banat, Sophorolipids production by Candida bombicola ATCC 22214 and its potential application in microbial enhanced oil recovery, Front. Microbiol. 6 (2015). [8] L. Ławniczak, R. Marecik, L. Chrzanowski, Contributions of biosurfactants to natural or induced bioremediation, Appl. Microbiol. Biotechnol. 97 (2013) 2327–2339. [9] A.M. Davila, R. Marchal, J.P. Vandecasteele, Sophorose lipid fermentation with differentiated substrate supply for growth and production phases, Appl. Microbiol. Biotechnol. 47 (1997) 496–501. [10] Institut Francais Du Petrole. Method of production of sophorosides by fermentation with fed batch supply of fatty acid esters or oils. US. Patent. US 5616479A. [11] U. Rau, S. Hammen, R. Heckmann, V. Wray, S. Lang, Sophorolipids A source for novel compounds, Ind. Crops Prod. 13 (2001) 85–92. [12] A.M. Davila, R. Marchal, J.P. Vandecasteele, Kinetics and balance of a fermentation free from product inhibition: sophorose lipid production by Candida bombicola, Appl. Microbiol. Biotechnol. 38 (1992) 6–11. [13] N. Kosaric, F.V. Sukan, Biosurfactants:Production and Utilization-Processes, Technologies and Economics, CRC Press, Boka Raton, 2014. [14] U. Rau, C. Manzke, F. Wagner, Influence of substrate supply on the production of sophorose lipids by Candida bombicola ATCC 22214, Biotechnol. Lett. 18 (1996) 149–154. [15] H.J. Daniel, R.T. Otto, M. Reuss, C. Syldatk, Sophorolipid production with high yields on whey concentrate and rapeseed oil without consumption of lactose, Biotechnol. Lett. 20 (1998) 805–807. [16] O. Stüwer, R. Hommel, D. Haferburg, H.P. Kleber, Production of crystalline surface-active glycolipids by a strain of torulopsis apicola, J. Biotechnol. 6 (1987) 259–269.
  • 10. Please cite this article in press as: B.M. Dolman, et al., Integrated sophorolipid production and gravity separation, Process Biochem (2016), http://dx.doi.org/10.1016/j.procbio.2016.12.021 ARTICLE IN PRESSG Model PRBI-10893; No.of Pages10 10 B.M. Dolman et al. / Process Biochemistry xxx (2016) xxx–xxx [17] H.J. Daniel, M. Reuss, C. Syldatk, Production of sophorolipids in high concentration from deproteinized whey and rapeseed oil in a two stage fed batch process using Candida bombicola ATCC 22214 and Cryptococcus curvatus ATCC 20509, Biotechnol. Lett. 20 (1998) 1153–1156. [18] G.L. Kleman, W.R. Strohl, High cell density and high-productivity microbial fermentation, Curr. Opin. Biotechnol. 3 (1992) 93–98. [19] Institut Francals du Petrole. Process for the production of sophorolipids by cyclic fermentation with feed of fatty acid esters or oils. US. Patent. US 5879913A. [20] V. Guilmanov, A. Ballistreri, G. Impallomeni, R.A. Gross, Oxygen transfer rate and sophorose lipid production by Candida bombicola, Biotechnol. Bioeng. 77 (2002) 489–494. [21] S. Alonso, P.J. Martin, Impact of foaming on surfactin production by Bacillus subtilis: implications on the development of integrated in situ foam fractionation removal systems, Biochem. Eng. J. 110 (2016) 125–133. [22] J. Beuker, A. Steier, A. Wittgens, F. Rosenau, M. Henkel, R. Hausmann, Integrated foam fractionation for heterologous rhamnolipid production with recombinant Pseudomonas putida in a bioreactor, AMB Express 6 (2016) 11. [23] J.B. Winterburn, A.B. Russell, P.J. Martin, Integrated recirculating foam fractionation for the continuous recovery of biosurfactant from fermenters, Biochem. Eng. J. 54 (2011) 132–139. [24] A. Daverey, K. Pakshirajan, Production, characterization, and properties of sophorolipids from the yeast Candida bombicola using a low-cost fermentative medium, Appl. Biochem. Biotechnol. 158 (2009) 663–674. [25] R. Gao, M. Falkeborg, X. Xu, Z. Guo, Production of sophorolipids with enhanced volumetric productivity by means of high cell density fermentation, Appl. Microbiol. Biotechnol. 97 (2013) 1103–1111. [26] Y.B. Kim, H.S. Yun, E.K. Kim, Enhanced sophorolipid production by feeding-rate-controlled fed-batch culture, Bioresour. Technol. 100 (2009) 6028–6032. [27] U. Rau, L.A. Nguyen, H. Roeper, H. Koch, S. Lang, Downstream processing of mannosylerythritol lipids produced by Pseudozyma aphidis, Eur. J. Lipid. Sci. Technol. 107 (2005) 373–380. [28] M.C. Cuellar, L.A.M. van der Wielen, Recent advances in the microbial production and recovery of apolar molecules, Curr. Opin. Biotechnol. 33 (2015) 39–45. [29] K. Joshi-Navare, P. Khanvilkar, A. Prabhune, Jatropha oil derived sophorolipids: production and characterization as laundry detergent additive, Biochem. Res. Int. 2013 (2013) 1–11.