Sequential Flash NanoPrecipitation For The Scalable Fo 2023 International Jo

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

International Journal of Pharmaceutics 640 (2023) 122985

Contents lists available at ScienceDirect

International Journal of Pharmaceutics


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

Sequential Flash NanoPrecipitation for the scalable formulation of stable


core-shell nanoparticles with core loadings up to 90%
Nicholas J. Caggiano a, 1, Satya K. Nayagam a, 1, Leon Z. Wang a, Brian K. Wilson a, Parker Lewis a,
Shadman Jahangir a, Rodney D. Priestley a, b, Robert K. Prud’homme a, Kurt D. Ristroph a, *
a
Department of Chemical and Biological Engineering, Princeton University, Princeton, NJ 08544, United States
b
Princeton Materials Institute, Princeton University, Princeton, NJ 08544, United States

A B S T R A C T

Flash NanoPrecipitation (FNP) is a scalable, single-step process that uses rapid mixing to prepare nanoparticles with a hydrophobic core and amphiphilic stabilizing
shell. Because the two steps of particle self-assembly – (1) core nucleation and growth and (2) adsorption of a stabilizing polymer onto the growing core surface –
occur simultaneously during FNP, nanoparticles formulated at core loadings above approximately 70% typically exhibit poor stability or do not form at all.
Additionally, a fundamental limit on the concentration of total solids that can be introduced into the FNP process has been reported previously. These limits are
believed to share a common mechanism: entrainment of the stabilizing polymer into the growing particle core, leading to destabilization and aggregation. Here, we
demonstrate a variation of FNP which separates the nucleation and stabilization steps of particle formation into separate sequential mixers. This scheme allows the
hydrophobic core to nucleate and grow in the first mixing chamber unimpeded by adsorption of the stabilizing polymer, which is later introduced to the growing
nuclei in the second mixer. Using this Sequential Flash NanoPrecipitation (SNaP) technique, we formulate stable nanoparticles with up to 90% core loading by mass
and at 6-fold higher total input solids concentrations than typically reported.

1. Introduction (kilograms) (Bobbala et al., 2018; Feng et al., 2019a; Li et al., 2014; Yu
et al., 2021; Zeng et al., 2019). Although FNP typically produces NPs
Nanoparticle (NP) delivery of hydrophobic compounds is of with core loadings of 33–70%, even higher drug loadings are desired to
increasing interest for solubility enhancement since a large proportion reduce the mass of excipients relative to the therapeutic compound
of new small molecule drugs exhibit poor solubility in water, thereby (Saad and Prud’homme, 2016; Shen et al., 2017; Tao et al., 2019; Zhu,
limiting oral bioavailability (Lipinski, 2000; Savjani et al., 2012; Thayer, 2014). Additionally, producing NPs at higher mass concentrations re­
2010). NP delivery is also of interest for parenteral and oral delivery of duces solvent waste and improves process mass intensity.
hydrophilic compounds, including biomolecules such as proteins, pep­ Rapid precipitation processes, including FNP, fail to produce stable
tides, and nucleic acids (Anselmo et al., 2019; Paunovska et al., 2022). nanoparticles when either (a) the total solids concentration during
Several approaches to prepare nanoparticles have been presented and precipitation, or (b) the concentration of core material relative to the
reviewed;(Khan et al., 2019; Kumar et al., 2020) however, many tech­ concentration of the stabilizer, is too high. Normally, the maximum
niques are limited by low core loading (mass fraction of core material achievable total solids concentration of NPs is approximately 0.1 wt% in
per mass of total NP) or poor process scalability (Cai et al., 2015; Qin the final dispersion (Pagels et al., 2018; Pustulka et al., 2013). At low
et al., 2017; Shen et al., 2017; Zhang et al., 2015). Our research has total solids concentrations, the ratio of core material to stabilizer (e.g.,
focused on a kinetically controlled, polymer-directed, rapid precipita­ drug loading) for nonionic hydrophobic species is usually limited to
tion process termed Flash NanoPrecipitation (FNP). FNP relies on 20–50% (Pagels et al., 2018). The limitations in total solids and drug
controlled turbulent micromixing in specially designed mixing cham­ loading are related to the mechanism of self-assembly. It has been sug­
bers to produce supersaturations as high as 10,000 in 1.5 ms (Johnson gested that for formulations with core loadings above 70%, the ability of
and Prud’homme, 2003a; Pinkerton et al., 2015). The technique is stabilizing polymer molecules to adsorb strongly onto a particle nucleus
scalable and has been demonstrated to produce the same monodisperse (so-called ‘anchoring efficiency’) is decreased due to the presence of
nanoparticles from laboratory scales (milligrams) to industry scales stabilizer molecules which have become kinetically trapped within the

* Corresponding author at: Department of Agricultural and Biological Engineering, Purdue University, West Lafayette, IN 47909, United States.
E-mail address: ristroph@purdue.edu (K.D. Ristroph).
1
Contributed equally.

https://doi.org/10.1016/j.ijpharm.2023.122985
Received 28 February 2023; Received in revised form 18 April 2023; Accepted 21 April 2023
Available online 29 April 2023
0378-5173/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
N.J. Caggiano et al. International Journal of Pharmaceutics 640 (2023) 122985

growing particle core (Pustulka et al., 2013). This reduces the effective colloids (Budijono et al., 2010; Ungun et al., 2009). In SNaP, the delay
amount of available stabilizer and introduces packing inefficiencies in time between the initial core precipitation and the subsequent applica­
the drug-loaded NP core. The result is larger, more polydisperse nano­ tion of the stabilizing layer is determined by the residence time in the
particles which exhibit poor size stability over time (Fig. 1A). tubing connecting the first and second chambers. This delay time is
Temporally separating the precipitation of the core materials from intentionally kept to a minimum to limit the extent of core aggregation
the adsorption of the stabilizing polymer surface material is a strategy to which occurs between the precipitation and coating steps.
overcome the limitations of loading. One method to achieve this sepa­
ration is through the use of mixed solvent systems to effect sequential 2. Materials and methods
separation of components. For example, Liu, Weitz, et al. demonstrated
a two-step batch nanoprecipitation process where two sequential ali­ 2.1. Reagents
quots of antisolvent were added to an initial ternary solvent mixture
containing core and polymer stabilizer. The first antisolvent addition Polymeric core material polycaprolactone diol (PCL, 2 kDa) was
caused precipitation of the core material, while the stabilizer remained obtained from Sigma-Aldrich (St. Louis, MO) and used without further
soluble. After the second addition core–shell nanoparticles were pro­ purification. Hydroxypropyl methylcellulose acetate succinate
duced with loadings of up to 59% (Han et al., 2020; Liu et al., 2020). A (HPMCAS; AQOAT® AS-HF) was generously provided by Shin-Etsu
similar effect can also be produced in a continuous process using Chemical Co. (Tokyo, Japan). Tetrahydrofuran (THF, HPLC grade)
microfluidics to introduce antisolvents sequentially for the core and was purchased from Fisher Chemical (Hampton, NH). Sodium hydroxide
stabilizing compounds (Liu et al., 2017). However, both cases require and sodium chloride were purchased from Sigma-Aldrich (St. Louis,
careful design of solvent systems to achieve sequential precipitation. MO). Deionized water (18.2 MΩ•cm) was prepared using a Thermo
Additionally, there are limitations to scaling microfluidics to large-scale Scientific Nanopure Diamond water system (Waltham, MA).
nanoparticle production. In microfluidics, rapid solvent diffusion is
achieved using the micron length scales of the microfluidic device; in
small geometries the flows are laminar. In contrast, during turbulent 2.2. Sequential mixing and nanoparticle characterization
mixing in FNP the small length scales for diffusion are achieved by using
turbulence to establish solvent-antisolvent striations at the Kolmogorov 2.2.1. Preparation of FNP control formulations.
length scale (Johnson and Prud’homme, 2003b). FNP has been applied Flash NanoPrecipitation was performed using a confined impinging
to various payloads and stabilizing polymer chemistries and is highly jet mixer (CIJ) as reported previously (Han et al., 2012; Markwalter
scalable (Feng et al., 2019a; Pinkerton et al., 2017; Pustulka et al., 2013; et al., 2019; Saad and Prud’homme, 2016). In all nanoparticle formu­
Yan et al., 2021; Yu et al., 2021; Zeng et al., 2019). lations studied here, PCL was the core material and HPMCAS was the
Our process, which we term Sequential Flash NanoPrecipitation stabilizing material. In brief, a solvent stream was prepared by dis­
(SNaP), is accomplished by the sequential arrangement of two mixing solving polycaprolactone (PCL, 2 kDa) in tetrahydrofuran at various
chambers. In the first chamber, the hydrophobic core material is concentrations. An aqueous antisolvent stream was prepared by dis­
precipitated by rapid mixing and antisolvent-induced supersaturation. solving HPMCAS free acid at various concentrations in water with
Before the hydrophobic cores can aggregate beyond the desired size 15 mM NaCl along with 1.08 equivalents of NaOH to ionize the succinyl
scale, they enter a second FNP mixing chamber where the stabilizing groups on the polymer. The solvent and antisolvent solutions (0.5 mL
polymer is introduced, which adsorbs onto the NP core surfaces each) were loaded into separate Luer-lip syringes, which were attached
(Fig. 1B). Conceptually, this is similar to the ability to introduce coating to a CIJ and rapidly depressed. The mixer effluent was immediately
layers onto nanoparticle surfaces via FNP, as we have demonstrated for collected in a 4 mL water reservoir. Formulation details for conventional
the coating of latex spheres(D’Addio et al., 2012) or inorganic oxide FNP formulations, including stream compositions, NP loadings, size, and
polydispersity index can be found in Table S1.

Fig. 1. Principles of conventional FNP and sequential SNaP processes. (A) FNP utilizes a one-step mixing process where aggregation and stabilization occur
simultaneously. (B) Proposed two-step mixing process where aggregation occurs followed by stabilization.

2
N.J. Caggiano et al. International Journal of Pharmaceutics 640 (2023) 122985

2.2.2. Preparation of SNaP formulations. was used here as a model core material because it behaves similarly to
For sequential nanoprecipitation, two CIJ mixers were used. Two nonionic, small-molecule hydrophobic drugs during the precipitation
2.5 mL glass Luer-lock syringes were attached to the first mixer, one process. In the first mixing step, the nanoparticle core material (PCL)
containing 1 mL of a THF solvent stream with dissolved PCL and the was dissolved in a water-miscible organic solvent (tetrahydrofuran) and
other containing 1 mL of 15 mM NaCl in water. The effluent port from then rapidly mixed against an aqueous antisolvent stream using a CIJ
this mixer was attached to one inlet on the second CIJ mixer by PEEK mixer. The antisolvent stream in FNP is typically deionized water; in this
tubing (described below). To the remaining port of the second mixer was work, a NaCl concentration of 7.5 mM was maintained in all mixing
attached a 5 mL glass Luer-lock syringe containing 2 mL of HPMCAS chambers to prevent self-stabilization of polycaprolactone cores via the
(with 1.08 equivalents of NaOH) at various concentrations in water with electrostatic surface charge that has been observed on hydrophobic
7.5 mM NaCl. All three syringes were depressed simultaneously, and the surfaces in water (Roger and Cabane, 2012; Zhang et al., 2012). Control
effluent from the second mixer was immediately collected in a 6 mL experiments performed with NaCl and without stabilizer showed that
water reservoir. Formulation details for SNaP formulations, including PCL did not self-stabilize and instead formed large aggregates.
stream compositions, NP loadings, size, and polydispersity index can be Following the initial precipitation step, the effluent from the first CIJ
found in Table S2. mixer was fed via tubing to the second CIJ mixer, where it was mixed
with a solution containing the nanoparticle surface stabilizing material.
2.2.3. Size characterization by dynamic light scattering. To avoid redissolution of the PCL cores formed in the first mixer, the
Following formulation, aliquots were removed from nanoparticle solvent concentration in the second mixer must not be increased above
dispersions and diluted in water to a NP concentration of 0.1 mg mL− 1 that of the first mixer effluent (50% water / 50% THF, v/v). Since the CIJ
for characterization. Size and polydispersity index (PDI) were measured mixer has only two inlets and requires nearly equivalent stream flow
by dynamic light scattering (DLS) using a Zetasizer Nano-ZS (Malvern rates to drive momentum annihilation (Saad and Prud’homme, 2016),
Instruments, Malvern, UK) (n = 3). Particle size is reported as Z-average the two-CIJ-mixer setup in Fig. 2 requires that the solution introduced in
hydrodynamic diameter obtained from intensity-weighted distributions. the second mixer be primarily aqueous.
Nanoparticle stability was monitored and evaluated using both visual Due to this requirement, a water-soluble stabilizing material must be
observation and DLS. used in a sequential nanoprecipitation scheme using two CIJ mixers. In
this work we used hydroxypropyl methylcellulose acetate succinate
(HPMCAS), a semisynthetic cellulosic polymer reported previously as
2.3. Equipment and mixing setups
appropriately amphiphilic for application as a stabilizer in FNP nano­
particle formulations (Caggiano et al., 2022; Feng et al., 2019b, 2018;
A mixing stand was designed in-house to allow for depression of
Ristroph et al., 2019). However, many of the polymers studied in the
multiple syringes for the SNaP process in making small-scale formula­
FNP literature are water-insoluble diblock copolymers that are not
tions, as shown in Figure S1. The stand base plate was modified to seat
appropriate for this application. To use these in the SNaP process, the
two confined impinging jet (CIJ) mixers and, later, one CIJ and one
second mixer must have more than two inlets, or else not require equal
multi-inlet vortex mixer (MIVM) (Liu et al., 2008; Markwalter et al.,
stream volumes. The multi-inlet vortex mixer (MIVM)(Liu et al., 2008;
2019). Detailed CAD drawings can be found in Figure S2. For SNaP, the
Markwalter et al., 2019) with four inlets, which was developed for FNP
outlet of the first mixer was connected to one of the inlets of the second
as an alternative to the CIJ geometry, meets these requirements and will
mixer with 250 mm of 0.030-inch inner diameter polyether ether ketone
be a focus of future work on SNaP. Initial efforts to use an MIVM as the
(PEEK) tubing. Syringes were fixed on the remaining inlets of both
second mixer of SNaP (Fig. S1D-E) to enable a diblock copolymer sta­
mixers. The depression plate was then manually pushed down to initiate
bilizer were attempted for a formulation with a vitamin E acetate core
mixing. The time required to depress syringes using the mixing stand
and poly(styrene)-b-poly(ethylene glycol) (PS-b-PEG) stabilizer at 83%
was measured as tinject = 0.35s using video footage of the process and
core loadings. The resulting nanoparticles had a Z-average diameter of
averaging across multiple trials. This corresponds to a volumetric flow
approx. 200 nm (Figure S3); however, further development for the CIJ-
rate of 2.3 mL s− 1. For SNaP, the total time between when a fluid
to-MIVM system will be reserved for future work.
element entered the mixing chamber of the first mixer (mixing chamber
volume ~ 0.2 mL) and the second mixer was 0.14 s.
3.2. Nanoparticle formation
3. Results and discussion
3.2.1. Governing variables for precipitation and self-assembly.
3.1. Mixing setup and rationale for sequential nanoprecipitation Pagels et al. showed that the size of NPs prepared by FNP is inde­
pendent of the core material used, provided that material is sufficiently
The sequential mixing process can be divided into two steps: pre­ hydrophobic to rapidly precipitate in the mixed solvent environment of
cipitation and stabilization (Fig. 2). Low molecular weight (2 kDa) PCL an FNP mixing chamber (log Poctanol/water > ~4.5) (Pagels et al., 2018).

Fig. 2. (A) Schematic of SNaP using two connected confined impinging jet (CIJ) mixers, or (B) a CIJ mixer and a multi-inlet vortex mixer. Core material is
precipitated in the first mixer using an antisolvent stream before stabilizer is added in the second mixer.

3
N.J. Caggiano et al. International Journal of Pharmaceutics 640 (2023) 122985

Their work also showed that, for a given core material and given sta­ here.
bilizer, the two variables that control NP size are the core loading and As core loading was increased, the Z-average diameter of the SNaP
total mass concentration (TMC) of the formulation. Core loading is NPs decreased slightly and the PDI increased slightly. Fig. 3B shows the
discussed in the introduction: NP size increases as core loading in­ intensity-weighted particle size distributions for the five SNaP formu­
creases, and NPs either rapidly destabilize or do not form altogether at lations from Fig. 3A. All six DLS traces exhibited at least slight deviation
core loadings above 75% for block copolymer stabilized nanoparticles. from monomodality which became more prominent with increased core
Pagels defined the TMC in terms of the total solids in the organic feed loading. This was reflected in increased PDI values.
stream to FNP (Pagels et al., 2018). Since in SNaP there are multiple feed
streams and mixers, we here define the TMC as the sum of the mass of 3.2.3. Total mass concentration during mixing.
the core material and the mass of the stabilizer in the system divided by The average size and polydispersity index (PDI) of formulations
the final solution volume including the collection reservoir (in practice, prepared by FNP or SNaP at a fixed core loading of 50% and various
the difference between this definition and that used in Pagels et al. is a TMCs are shown in Fig. 4. Above a TMC of 2 mg mL− 1, conventional FNP
factor of 10). Higher TMC during the formulation process yields a more produced aggregates along with NPs, whereas SNaP produced NPs
concentrated final NP suspension, which translates to a reduction of without aggregates at TMCs up to 6 mg mL− 1, representing a six-fold
post-processing times (e.g., spray drying times or ultrafiltration con­ increase in the solids concentration in the effluent from the two-mixer
centration times) and an improvement of the overall process mass in­ setup. Again, close visual scrutiny was necessary to identify aggregates
tensity. For formulations of equal core loading, increasing the TMC in conventional FNP formulations at this scale; DLS measurements alone
increases the resulting NP size. Above a critical TMC, visible aggregates were not sufficient to determine whether aggregation occurred
are observed, indicating that not all core material is encapsulated. As (Figure S4).
with the upper limit of core loading, the likely mechanism for this
limitation is the entrainment of stabilizer molecules within NP cores. As 3.2.4. Particle size.
TMC increases, the average distance between two adjacent molecules For formulations in which both methods yielded NPs without ag­
during NP assembly decreases, which can lead to stabilizer molecules gregates in Figs. 3 and 4, NPs made by sequential precipitation were
becoming trapped among rapidly assembling core molecules. Typical consistently larger and had higher PDI values than those made by single-
TMCs for FNP formulations are around 1–2 mg mL− 1 (Markwalter et al., step precipitation. This was in line with expectations, as NP cores in the
2019). sequential precipitation process had more time to aggregate before
growth was arrested by HPMCAS adsorption in the second mixer. As
3.2.2. Nanoparticle core loading. described below, however, the delay time between mixers does not fully
The average size and polydispersity index (PDI) of formulations account for this difference in size.
prepared by FNP or SNaP at a fixed TMC of 1 mg mL− 1 and different core Pagels modified Smoluchowski’s model for diffusion-limited aggre­
loadings are given in Fig. 3A. Conventional FNP produced a combination gation to model NP assembly under FNP conditions (Pagels et al., 2018).
of NPs and visible aggregates at core loadings of 75% and higher. In One equation from the modified model can be used to back-calculate the
contrast, no visible aggregates were observed in SNaP formulations at amount of time required for NPs to assemble:
core loadings of 50%, 63%, 75%, 83%, and 90%. At 95% core loading, ( )13
some replicates exhibited aggregates while others did not. At 99% core r=
tkB Tccore
loading aggregates were consistently observed. Therefore, the maximum πμρcore
achievable core loading for this system is likely in the range of 90–95%. where NP radius r (meters) may be written as a function of time t
It is important to note that DLS measurements did not always detect the (seconds), absolute temperature T, concentration of core material ccore
aggregates, which tended to adhere to vial walls or form a film on the (kg m− 3) solution viscosity μ (kg m− 1 s− 1), core material density ρcore (kg
surface of the liquid in a vial. Close visual scrutiny was necessary to m− 3), and the Boltzmann constant kB (m2 kg s− 2 K− 1).
identify aggregates at this small scale (5–10 mL per formulation), and This model was used to calculate the difference in expected assembly
representative photographs are provided in the SI (Figure S4). DLS time between the two comparable formulations at 62.5% core loading in
measurements for the conventional FNP formulations produced at Fig. 3. These calculations estimate that the 196 nm diameter particles
loadings above 75% indicated that the population of non-aggregated produced by conventional FNP assembled over approximately 520 ms,
NPs had smaller average diameter and lower PDI than corresponding compared with 2,700 ms for the 340 nm diameter SNaP NPs. As the
SNaP NPs. However, these DLS measurements are not representative of delay time between the entrance from one mixer to the next mixers was
the true particle size distribution inclusive of the population of aggre­ estimated at only 140 ms, this delay alone does not appear to fully ac­
gates. At pharmaceutical scale, the formation of aggregates would lead count for the 2,200 ms difference in predicted assembly time.
to product losses and reduced reproducibility, hindering scale-up efforts Importantly, the residence time in a CIJ under standard operation is
(Feng et al., 2019a; Li et al., 2014; Zeng et al., 2019). As such, DLS only about 90 ms. For NP formulations with elevated core loadings or
measurements for visibly aggregated NP formulations are not reported high TMCs, then, the model predicts that particle assembly is not

Fig. 3. (A) Z-average size (circles) and polydispersity index (PDI, crosses) of NPs produced by SNaP (black) or conventional FNP (orange) plotted versus core loading
for a fixed TMC of 1 mg mL− 1, (B) DLS traces for the six SNaP formulations shown in (A).

4
N.J. Caggiano et al. International Journal of Pharmaceutics 640 (2023) 122985

Fig. 4. (A) Z-average size (circles) and polydispersity index (PDI, crosses) of NPs produced by SNaP (black) or conventional FNP (orange) plotted versus TMC for a
fixed core loading of 50%, (B) DLS traces for the four SNaP formulations shown in (A).

complete by the time a fluid element exits the final mixer of either sequential precipitation process.
conventional FNP or SNaP. Here, the mixer effluent containing still- When two CIJ mixers are used, the choice of amphiphilic stabilizer is
assembling NPs was collected in a downstream quench reservoir, limited to water-soluble species that can be introduced in an aqueous
where mixing was not as finely controlled as the rapid turbulent mixing stream into the second mixer. Moving forward, we recommend that the
inside the CIJ. This may have contributed to the larger sizes and higher second mixer be an MIVM for improved scalability and flexibility, as this
PDI values of the final SNaP NPs. Despite the difference in final size would result in better in-mixer and downstream solvent quality control
(discussed further below), the ‘pre-templating’ of the SNaP cores ap­ and would enable the use of a wider range of amphiphilic stabilizer
pears to have enabled stable particle formation at high core loadings and chemistries, including those that are only soluble in organic solvents. A
TMCs without the formation of visible aggregates. The 140 ms of time formulation prepared using this CIJ-to-MIVM setup using a vitamin E
for PCL cores to nucleate in SNaP prior to HPMCAS exposure in the core and block copolymer PS-b-PEG stabilizer was developed and is
second mixer should have been sufficient for nuclei 125 nm in diameter reported in SI (Figure S3). However, manual operation of a CIJ-to-
to form. These cores then began to be coated by HPMCAS in the second MIVM SNaP system is challenging and is limited by available syringe
mixer. volumes and geometries. Independent control of stream flowrates (and
If both formulations finished assembling after (poorly-controlled) thus solvent quality in the MIVM) can only be achieved using pumps to
dilution into additional water downstream of the good mixing – dilution feed each stream into the mixers. Development of a lab setup for CIJ-to-
which lowered the TMC and affected solvent quality (and therefore MIVM and MIVM-to-MIVM operation is currently being pursued in our
supersaturation) for this final stage of assembly – the NP cores made by laboratory.
conventional FNP may have been smaller because of longer exposure to For formulations where NPs can be formed by both SNaP and con­
the HPMCAS stabilizer. NP cores formed by conventional FNP were ventional FNP, SNaP NPs are larger in diameter. We attribute some of
exposed to HPMCAS for all of the pre-quench assembly time, whereas this difference to NP assembly which occurs downstream of the final CIJ
cores formed by SNaP were only exposed to HPMCAS for less than half of mixer in a less-controlled solvent environment. This would be solved by
the pre-quench assembly time. the CIJ-to-MIVM setup described above. However, it is conceivable that
The above example is an important point that should be considered increasing the delay time between mixers could lead to tunable particle
when preparing NPs by FNP using a CIJ mixer. For NPs that are expected size in the 300–1000 nm range, which is currently difficult to achieve by
to take much longer than 90 ms to assemble (predictable as a function of FNP and which may have applications in depot delivery or microparticle
TMC, core loading, and stabilizer), some or most particle assembly takes formulation. This will be another focus of future studies.
place downstream of the mixer. In industrial applications, the MIVM Funding sources
geometry removes the need for a downstream quench reservoir and is, Funding support has come from the Helen S Hunt Fund to the
therefore, recommended for formulations (whether SNaP or conven­ Princeton University School of Engineering and Science. We also
tional FNP) made at high TMCs and/or high core loadings, which are acknowledge support from the Bill and Melinda Gates Foundation under
expected to take more time to assemble than the residence time in the grants INV-010674, INV-010674, and INV-042629.
mixer.

3.2.5. Nanoparticle size stability Declaration of Competing Interest


All formulations made by SNaP in Figs. 3 and 4 remained stable in a
90:10 volumetric mixture of water to THF for 24 h (Figure S5). Size was The authors declare that they have no known competing financial
monitored by DLS, and vials were visually scrutinized for aggregate interests or personal relationships that could have appeared to influence
formation. Representative photographs from visual inspection of various the work reported in this paper.
formulation outcomes are included as Figure S4.
Appendix A. Supplementary material
4. Conclusions
Supplementary data to this article can be found online at https://doi.
The results presented here offer proof of principle for an extension of org/10.1016/j.ijpharm.2023.122985.
FNP technology that separates the nucleation and stabilization steps of
nanoparticle assembly. Using two sequentially connected FNP mixers, References
nanoparticles can be formulated at higher core loadings and TMCs than
in conventional single-step FNP. NPs with core loadings up to 90% were Anselmo, A.C., Gokarn, Y., Mitragotri, S., 2019. Non-invasive delivery strategies for
biologics. Nat. Rev. Drug Discov. 18, 19–40. https://doi.org/10.1038/nrd.2018.183.
formulated without the aggregates that form via conventional FNP at the Bobbala, S., Allen, S.D., Scott, E.A., 2018. Flash nanoprecipitation permits versatile
same mass loading and TMC. This improvement in loading and TMC are assembly and loading of polymeric bicontinuous cubic nanospheres. Nanoscale 10,
major advantages for this SNaP technique. Additionally, only minimal 5078–5088. https://doi.org/10.1039/C7NR06779H.
Budijono, S.J., Shan, J., Yao, N., Miura, Y., Hoye, T., Austin, R.H., Ju, Y., Prud’homme, R.
modification of existing FNP equipment is required to facilitate the K., 2010. Synthesis of stable block-copolymer-protected NaYF 4: Yb 3+, Er 3+ up-

5
N.J. Caggiano et al. International Journal of Pharmaceutics 640 (2023) 122985

converting phosphor nanoparticles. Chem. Mater. 22, 311–318. https://doi.org/ Pagels, R.F., Edelstein, J., Tang, C., Prud’homme, R.K., 2018. Controlling and predicting
10.1021/cm902478a. nanoparticle formation by block copolymer directed rapid precipitations. Nano Lett.
Caggiano, N.J., Wilson, B.K., Priestley, R.D., Prud’homme, R.K., 2022. Development of 18, 1139–1144. https://doi.org/10.1021/acs.nanolett.7b04674.
an in vitro release assay for low-density cannabidiol nanoparticles prepared by Flash Paunovska, K., Loughrey, D., Dahlman, J.E., 2022. Drug delivery systems for RNA
NanoPrecipitation. Mol. Pharm. 19, 1515–1525. https://doi.org/10.1021/acs. therapeutics. Nat. Rev. Genet. 23, 265–280. https://doi.org/10.1038/s41576-021-
molpharmaceut.2c00041. 00439-4.
Cai, K., He, X., Song, Z., Yin, Q., Zhang, Y., Uckun, F.M., Jiang, C., Cheng, J., 2015. Pinkerton, N.M., Gindy, M.E., Calero-DdelC, V.L., Wolfson, T., Pagels, R.F., Adler, D.,
Dimeric drug polymeric nanoparticles with exceptionally high drug loading and Gao, D., Li, S., Wang, R., Zevon, M., Yao, N., Pacheco, C., Therien, M.J., Rinaldi, C.,
quantitative loading efficiency. J. Am. Chem. Soc. 137, 3458–3461. https://doi.org/ Sinko, P.J., Prud’homme, R.K., 2015. Single-step assembly of multimodal imaging
10.1021/ja513034e. nanocarriers: MRI and long-wavelength fluorescence imaging. Adv. Healthc. Mater.
D’Addio, S.M., Saad, W., Ansell, S.M., Squiers, J.J., Adamson, D.H., Herrera-Alonso, M., 4, 1376–1385. https://doi.org/10.1002/adhm.201400766.
Wohl, A.R., Hoye, T.R., Macosko, C.W., Mayer, L.D., Vauthier, C., Prud’homme, R.K., Pinkerton, N.M., Behar, L., Hadri, K., Amouroux, B., Mingotaud, C., Talham, D.R.,
2012. Effects of block copolymer properties on nanocarrier protection from in vivo Chassaing, S., Marty, J.-D., 2017. Ionic Flash NanoPrecipitation (iFNP) for the facile,
clearance. J. Control. Release 162, 208–217. https://doi.org/10.1016/j. one-step synthesis of inorganic–organic hybrid nanoparticles in water. Nanoscale 9,
jconrel.2012.06.020. 1403–1408. https://doi.org/10.1039/C6NR09364G.
Feng, J., Zhang, Y., McManus, S.A., Ristroph, K.D., Lu, H.D., Gong, K., White, C.E., Pustulka, K.M., Wohl, A.R., Lee, H.S., Michel, A.R., Han, J., Hoye, T.R., McCormick, A.V.,
Prud’homme, R.K., 2018. Rapid recovery of clofazimine-loaded nanoparticles with Panyam, J., Macosko, C.W., 2013. Flash Nanoprecipitation: particle structure and
long-term storage stability as anti- cryptosporidium therapy. ACS Appl. Nano Mater. stability. Mol. Pharm. 10, 4367–4377. https://doi.org/10.1021/mp400337f.
1, 2184–2194. https://doi.org/10.1021/acsanm.8b00234. Qin, S.-Y., Zhang, A.-Q., Cheng, S.-X., Rong, L., Zhang, X.-Z., 2017. Drug self-delivery
Feng, J., Markwalter, C.E., Tian, C., Armstrong, M., Prud’homme, R.K., 2019a. systems for cancer therapy. Biomaterials 112, 234–247. https://doi.org/10.1016/j.
Translational formulation of nanoparticle therapeutics from laboratory discovery to biomaterials.2016.10.016.
clinical scale. J. Transl. Med. 17, 200. https://doi.org/10.1186/s12967-019-1945-9. Ristroph, K.D., Feng, J., McManus, S.A., Zhang, Y., Gong, K., Ramachandruni, H.,
Feng, J., Zhang, Y., McManus, S.A., Qian, R., Ristroph, K.D., Ramachandruni, H., White, C.E., Prud’homme, R.K., 2019. Spray drying OZ439 nanoparticles to form
Gong, K., White, C.E., Rawal, A., Prud’homme, R.K., 2019b. Amorphous stable, water-dispersible powders for oral malaria therapy. J. Transl. Med. 17, 97.
nanoparticles by self-assembly: processing for controlled release of hydrophobic https://doi.org/10.1186/s12967-019-1849-8.
molecules. Soft Matter 15, 2400–2410. https://doi.org/10.1039/C8SM02418A. Roger, K., Cabane, B., 2012. Why are hydrophobic/water interfaces negatively charged?
Han, F.Y., Liu, Y., Kumar, V., Xu, W., Yang, G., Zhao, C.-X., Woodruff, T.M., Whittaker, A. Angew. Chem. Int. Ed. 51, 5625–5628. https://doi.org/10.1002/anie.201108228.
K., Smith, M.T., 2020. Sustained-release ketamine-loaded nanoparticles fabricated Saad, W.S., Prud’homme, R.K., 2016. Principles of nanoparticle formation by flash
by sequential nanoprecipitation. Int. J. Pharm. 581, 119291 https://doi.org/ nanoprecipitation. Nano Today 11, 212–227. https://doi.org/10.1016/j.
10.1016/j.ijpharm.2020.119291. nantod.2016.04.006.
Han, J., Zhu, Z., Qian, H., Wohl, A.R., Beaman, C.J., Hoye, T.R., Macosko, C.W., 2012. Savjani, K.T., Gajjar, A.K., Savjani, J.K., 2012. Drug solubility: importance and
A simple confined impingement jets mixer for flash nanoprecipitation. J. Pharm. Sci. enhancement techniques. ISRN Pharm. 2012, 1–10. https://doi.org/10.5402/2012/
101, 4018–4023. https://doi.org/10.1002/jps.23259. 195727.
Johnson, B.K., Prud’homme, R.K., 2003a. Mechanism for rapid self-assembly of block Shen, S., Wu, Y., Liu, Y., Wu, D., 2017. High drug-loading nanomedicines: progress,
copolymer nanoparticles. Phys. Rev. Lett. 91, 118302 https://doi.org/10.1103/ current status, and prospects. Int. J. Nanomedicine 12, 4085–4109. https://doi.org/
PhysRevLett.91.118302. 10.2147/IJN.S132780.
Johnson, B.K., Prud’homme, R.K., 2003b. Chemical processing and micromixing in Tao, J., Chow, S.F., Zheng, Y., 2019. Application of flash nanoprecipitation to fabricate
confined impinging jets. AIChE J. 49, 2264–2282. https://doi.org/10.1002/ poorly water-soluble drug nanoparticles. Acta Pharm. Sin B, SI: Enhancement of
aic.690490905. dissolution and oral bioavailability of poorly water-soluble drugs 9, 4–18. https://
Khan, I., Saeed, K., Khan, I., 2019. Nanoparticles: Properties, applications and toxicities. doi.org/10.1016/j.apsb.2018.11.001.
Arab. J. Chem. 12, 908–931. https://doi.org/10.1016/j.arabjc.2017.05.011. Thayer, A.M., 2010. Finding solutions: custom manufacturers take on drug solubility
Kumar, R., Dalvi, S.V., Siril, P.F., 2020. Nanoparticle-based drugs and formulations: issues to help pharmaceutical firms move products through development. Chem.
current status and emerging applications. ACS Appl. Nano Mater. 3, 4944–4961. Eng. News Arch. 88, 13–18. https://doi.org/10.1021/cen-v088n022.p013.
https://doi.org/10.1021/acsanm.0c00606. Ungun, B., Prud’homme, R.K., Budijon, S.J., Shan, J., Lim, S.F., Ju, Y., Austin, R., 2009.
Li, K.-K., Zhang, X., Huang, Q., Yin, S.-W., Yang, X.-Q., Wen, Q.-B., Tang, C.-H., Lai, F.-R., Nanofabricated upconversion nanoparticles for photodynamic therapy. Opt. Exp. 17
2014. Continuous preparation of zein colloidal particles by Flash NanoPrecipitation (1), 80. https://doi.org/10.1364/OE.17.000080.
(FNP). J. Food Eng. 127, 103–110. https://doi.org/10.1016/j.jfoodeng.2013.12.001. Yan, X., Bernard, J., Ganachaud, F., 2021. Nanoprecipitation as a simple and
Lipinski, C.A., 2000. Drug-like properties and the causes of poor solubility and poor straightforward process to create complex polymeric colloidal morphologies. Adv.
permeability. J. Pharmacol. Toxicol. Methods. 44 (1), 235–249. https://doi.org/ Colloid Interface Sci. 294, 102474 https://doi.org/10.1016/j.cis.2021.102474.
10.1016/s1056-8719(00)00107-6. Yu, M., Zhang, W., Guo, Z., Wu, Y., Zhu, W., 2021. Engineering nanoparticulate organic
Liu, Y., Cheng, C., Liu, Y., Prud’homme, R.K., Fox, R.O., 2008. Mixing in a multi-inlet photocatalysts via a scalable flash nanoprecipitation process for efficient hydrogen
vortex mixer (MIVM) for flash nano-precipitation. Chem. Eng. Sci. 63 (11), production. Angew. Chem. Int. Ed. 60, 15590–15597. https://doi.org/10.1002/
2829–2842. https://doi.org/10.1016/j.ces.2007.10.020. anie.202104233.
Liu, Y., Yang, G., Baby, T., Tengjisi, Chen, D., Weitz, D.A., Zhao, C.-X., 2020. Stable Zeng, Z., Dong, C., Zhao, P., Liu, Z., Liu, L., Mao, H.-Q., Leong, K.W., Gao, X., Chen, Y.,
polymer nanoparticles with exceptionally high drug loading by sequential 2019. Scalable production of therapeutic protein nanoparticles using flash
nanoprecipitation. Angew. Chem. 59 (12), 4720–4728. https://doi.org/10.1002/ nanoprecipitation. Adv. Healthc. Mater. 8 (6), e1801010. https://doi.org/10.1002/
anie.201913539. adhm.201801010.
Liu, D., Zhang, H., Cito, S., Fan, J., Mäkilä, E., Salonen, J., Hirvonen, J., Sikanen, T.M., Zhang, J., Li, S., An, F.-F., Liu, J., Jin, S., Zhang, J.-C., Wang, P.C., Zhang, X., Lee, C.-S.,
Weitz, D.A., Santos, H.A., 2017. Core/shell nanocomposites produced by superfast Liang, X.-J., 2015. Self-carried Curcumin Nanoparticles for In vitro and in vivo
sequential microfluidic nanoprecipitation. Nano Lett. 17, 606–614. https://doi.org/ cancer therapy with real-time monitoring of drug release. Nanoscale 7,
10.1021/acs.nanolett.6b03251. 13503–13510. https://doi.org/10.1039/c5nr03259h.
Markwalter, C.E., Pagels, R.F., Wilson, B.K., Ristroph, K.D., Prud’homme, R.K., 2019. Zhang, C., Pansare, V.J., Prud’homme, R.K., Priestley, R.D., 2012. Flash
Flash NanoPrecipitation for the encapsulation of hydrophobic and hydrophilic nanoprecipitation of polystyrene nanoparticles. Soft Matter 8 (1), 86–93. https://
compounds in polymeric nanoparticles. J. Vis. Exp. 58757 https://doi.org/10.3791/ doi.org/10.1039/C1SM06182H.
58757. Zhu, Z., 2014. Flash nanoprecipitation: prediction and enhancement of particle stability
via drug structure. Mol. Pharm. 11, 776–786. https://doi.org/10.1021/mp500025e.

You might also like