ảnh hưởng của thống số dung dịch

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

Subscriber access provided by UNIVERSITY OF LEEDS

Review
A review of the fundamental principles
and applications of solution blow spinning
John L Daristotle, Adam M. Behrens, Anthony D. Sandler, and Peter Kofinas
ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.6b12994 • Publication Date (Web): 01 Dec 2016
Downloaded from http://pubs.acs.org on December 4, 2016

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society.
1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
A review of the fundamental principles and
9
10
11
12 applications of solution blow spinning
13
14
15
16
17 John L. Daristotle◊, Adam M. Behrens◊, Anthony D. Sandler×, Peter Kofinas*◊
18
19
20 Fischell Department of Bioengineering◊
21
22
23 2330 Jeong H. Kim Engineering Building
24
25 University of Maryland
26
27
28
College Park, MD 20742, USA
29
30 E-mail: kofinas@umd.edu
31
32
33 Sheikh Zayed Institute for Pediatric Surgical Innovation×
34
35
36 Joseph E. Robert Jr. Center for Surgical Care
37
38 Children’s National Medical Center
39
40
41
111 Michigan Avenue, NW Washington, DC 20010, USA
42
43
44 Keywords: solution blow spinning, nanofibers, nonwovens, fiber spinning, polymers,
45
46 nanocomposites, microfibers
47
48
49
50
51
52
53 Solution blow spinning (SBS) is a technique that can be used to deposit fibers in situ at low cost
54
55 for a variety of applications, which include biomedical materials and flexible electronics. This
56
57 review is intended to provide an overview of the basic principles and applications of SBS. We
58
59
60
ACS Paragon Plus Environment
1
ACS Applied Materials & Interfaces Page 2 of 55

1
2
3
first describe a method for creating a spinnable polymer solution and stable polymer solution jet
4
5
6 by manipulating parameters such as polymer concentration and gas pressure. This method is
7
8 based on fundamental insights, theoretical models, and empirical studies. We then discuss the
9
10
11
unique bundled morphology and mechanical properties of fiber mats produced by SBS, and how
12
13 they compare with electrospun fiber mats. Applications of SBS in biomedical engineering are
14
15 highlighted, showing enhanced cell infiltration and proliferation versus electrospun fiber
16
17
18 scaffolds and in situ deposition of biodegradable polymers. We also discuss the impact of SBS in
19
20 applications involving textiles and electronics, including ceramic fibers and conductive
21
22 composite materials. Strategies for future research are presented that take advantage of direct and
23
24
25 rapid polymer deposition via cost-effective methods.
26
27
28
29
30
31
32 1. INTRODUCTION
33
34
35 Solution blow spinning (SBS) is a fiber fabrication process that requires two parallel concentric
36
37
38 fluid streams: a polymer dissolved in a volatile solvent and a pressurized gas that flows around
39
40 the polymer solution, creating fibers that are deposited in the direction of gas flow (Figure 1A).
41
42 Generally, a SBS setup consists of a compressed gas source for delivering the carrier gas and a
43
44
45 syringe pump for polymer solution (Figure 1B). The two streams can be easily integrated into a
46
47 simple, easy to manufacture device (Figure 1C) or generated by using a commercially-available
48
49
airbrush (Figure 1D).1,2 SBS also has the ability to deposit conformal fibers onto both planar and
50
51
52 non-planar substrates with a deposition rate that is approximately ten times faster than
53
54 conventional electrospinning.2–4 The unique attributes of SBS provide a means to explore the
55
56
57
utility of non-woven fibrous materials in new fields. Such materials have been investigated for
58
59
60
ACS Paragon Plus Environment
2
Page 3 of 55 ACS Applied Materials & Interfaces

1
2
3
use in a wide variety of applications including electronics, filtration, and tissue engineering. New
4
5
6 opportunities for applications include rapid biological scaffold generation and custom in situ
7
8 materials fabrication.
9
10
11
12 In SBS, the carrier solvent evaporates quickly before the polymer fibers deposit on the
13
14 collection surface. Although acute exposure to high concentrations of a solvent such as acetone
15
16 may be toxic, studies have shown that SBS from acetone directly onto cells did not affect
17
18
19 viability.5 This allows SBS to be a biocompatible process. SBS does not require an electric field
20
21 and uses a simple apparatus. The method is therefore easily implemented using inexpensive,
22
23 transportable, and handheld equipment.2 When deployed as a technique to deposit materials in
24
25
26 situ, SBS is capable of additive deposition of microfibers or nanofibers in custom conformal
27
28 geometries.
29
30
31
32
In comparison to one of the most widely employed fiber fabrication techniques,
33
34 electrospinning, SBS has fewer process requirements and variables. In electrospinning, polymers
35
36 are commonly dissolved in highly toxic chlorinated or fluorinated solvents such as
37
38
39 dichloromethane, trifluoroethanol, or hexafluoro-2-propanol, which produce narrower, more
40
41 consistent fibers because of their relatively high dielectric constants.6,7 An additional limitation
42
43 of electrospinning is that a large electric field must be applied to facilitate fiber production.8
44
45
46 Fiber production is typically slow but well-controlled, precisely depositing a narrow distribution
47
48 of fibers with solution deposition rates on the order of 1 ml hr–1.2 These factors limit both the
49
50
commercial applicability of electrospun fibers and the capability of rapidly applying fibers for an
51
52
53 immediate in situ indication.9
54
55
56
57
58
59
60
ACS Paragon Plus Environment
3
ACS Applied Materials & Interfaces Page 4 of 55

1
2
3
SBS uses processes similar to those used in industrial methods of fiber production, which
4
5
6 enable its future applicability in large scale production. For example, in a melt spinning or dry
7
8 spinning process, a gas is used to cool or evaporate solvent from fibers, respectively, after
9
10
11
extrusion through a spinneret by pumping. SBS uses a pressurized gas to both drive extrusion of
12
13 the polymer solution and cause solvent evaporation, creating a polymer fiber in a simpler one-
14
15 step process. Industrial production techniques allow for the continuous mass production of long
16
17
18 fibers used in textile manufacturing applications. SBS allows for scalable production and
19
20 deposition at the point and site of use. A compact deposition device can be easily manipulated by
21
22 hand to deposit fibers in only the target area. Because the process is contained within the
23
24
25 spinning device and the solvent evaporates within the working distance (typically 10–20 cm, but
26
27 varies based on solvent volatility), polymer fibers are deposited with no further drying, cooling,
28
29
or washing necessary.
30
31
32
33 Alternatives to electrospinning and common industrial spinning methods have been developed
34
35 and characterized for specialty applications. These include techniques such as gas jet spinning,
36
37
38 nozzle-free centrifugal spinning, rotary jet spinning, and flash-spinning.10–14 Reviews on fiber
39
40 spinning techniques have included brief introductions to SBS or similar techniques.9,15 An
41
42 increasing number of papers have been published on the subject (Figure 2), reflecting growing
43
44
45 interest in this technology. Depending on the field, SBS research has been denoted “solution
46
47 blowing” or “airbrushing”, but these techniques share the same governing principles. This review
48
49
will compile research on techniques with similar principles, offering a comprehensive overview
50
51
52 of SBS while highlighting recent applications and future value to researchers.
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
4
Page 5 of 55 ACS Applied Materials & Interfaces

1
2
3
2. BASIC PRINCIPLES OF FIBER SPINNING
4
5
6
7 SBS fiber generation depends on the molecular weight of the polymer, concentration, and
8
9 viscosity of the polymer solution, and process variables such as gas pressure and polymer
10
11
12 solution flow rate. These parameters have direct influence on critical characteristics that enable
13
14 the formation of a fiber-producing jet of polymer solution. The relationships between these
15
16 variables and fiber morphology and diameter have been empirically and theoretically studied.
17
18
19
20 2.1. Theoretical Background and Modeling. The ability to form fibers from a jet of polymer
21
22 solution is primarily governed by the entanglement of polymer chains. The overlap concentration
23
24
(c*) represents the critical point when polymer coils in solution begin to overlap, causing
25
26
27 entanglements. When the overlap concentration is reached, a polymer solution becomes semi-
28
29 dilute, and the interaction between entangled polymer chains causes an increase in viscosity. c*
30
31
32
is best estimated by an equation for the semi-dilute, good solvent polymer environment:16
33
34
 
35  ∗ = 6  /(8 〈  〉 ) (1)
36
37
An approximation of the mean-square end-to-end distance 〈  〉 for the polymer-solvent system
38
39
40 is necessary to calculate the overlap concentration required to form fibers (such as Equation 2).17
41
42
43 
〈  〉 =       (2)
44 
45
46 This can be found from the Flory expansion factor α, the characteristic ratio C∞, and bond length
47
48
l, which are used to represent deviations from ideality of polymer chain dimensions.18 The
49
50
51 weight-average molecular weight (Mw) of the polymer of interest and Avogadro’s number (Na) is
52
53 also required.
54
55
56
57
58
59
60
ACS Paragon Plus Environment
5
ACS Applied Materials & Interfaces Page 6 of 55

1
2
3
Srinivasan et al. previously used the overlap concentration to explain fiber formation in an
4
5
6 SBS process.19 It has also been used to determine the onset of fiber formation in an
7
8 electrospinning process, which shares the parameters of polymer concentration, polymer solution
9
10
11
viscosity, and polymer solution flow rate but not electric field strength or electric field
12
13 geometry.20 The overlap concentration (c*) represents the critical point when polymer
14
15 entanglement becomes significant enough to stabilize the polymer jet, overcoming the inertio-
16
17
18 capillary forces that drive bead formation instead of fiber formation under dilute conditions
19
20 (Figure 3A). It has also been proposed that the entanglement concentration, Ce ≈ 10c*,
21
22 approximates the concentration above which there will be no bead formation at all.21
23
24
25
26 The interplay between the viscous forces, inertio-capillary forces, and polymer relaxation time,
27
28 which govern fiber formation of a polymer solution jet, can be represented by two dimensionless
29
30
31
numbers: Deborah number (De, Equation 3), and Ohnesorge number (Oh, Equation 4).22
32
33
34  = / !"#$ /% (3)
35
36
&ℎ = (# / !%"# (4)
37
38
39 The exponential increase in zero-shear viscosity (η0) caused by exceeding the overlap
40
41 concentration causes viscous forces to dominate inertio-capillary forces, leading to fiber
42
43 formation as observed in these studies. However, it is also evident from De that inertio-capillary
44
45
46 forces must also be overcome in part by polymer relaxation time (λ), the time scale for
47
48 viscoelastic behavior of a polymer solution. λ increases proportionally to the entanglement of
49
50 polymer chains, and is insignificant for polymers with Mw less than the entanglement molecular
51
52
53 weight, Me.23 This indicates that a polymer must have a minimum molecular weight, Me, to make
54
55 a solution capable of forming fibers. Inertio-capillary forces, represented by the denominator of
56
57
58
59
60
ACS Paragon Plus Environment
6
Page 7 of 55 ACS Applied Materials & Interfaces

1
2
3
both dimensionless numbers, can be altered by manipulating the solution density (ρ), solution
4
5
6 surface tension (σ), and nozzle radius (r0).
7
8
9 Investigations into the influence of molecular weight on fiber formation in fiber spinning
10
11
12 process have yielded two additional insights: First, high molecular weight polymers can be spun
13
14 into fibers at concentrations lower than their overlap concentration.19 Srinivasan et al. confirmed
15
16 the significance of De to fiber formation by estimating it for various polymer solutions over a
17
18
19 range of polymer molecular weights and concentrations, all of which were at or below c*.
20
21 Despite Oh being similar for all solutions, increasing poly(methyl methacrylate) molecular
22
23 weight produced increasing λ, and therefore increasing De. This increase in polymer relaxation
24
25
26 time coincided with the observation of increasingly fiber-like morphology, culminating with
27
28 fiber formation at the highest molecular weight polymer solution, 2200 kDa, which possessed a
29
30
31
concentration c/c* = 0.72. Second, the extensibility average molecular weight (Equation 5) can
32
33 be used with the excluded volume constant v to estimate a minimum spinning concentration for
34
35 polymer blends (Equation 6):21
36
37
38 0
39 ∑- ,- -./ /.0 0

40 M* =  ∑- ,- -
 = (∑2 12 2345 )/.0 (5)
41
42
6728 ~ *:(;43) (6)
43
44
45 SBS has been modeled as a Newtonian thin liquid jet moving in air.24 Choosing a Newtonian
46
47 model implies that the viscous forces dominate inertio-capillary forces, i.e. the polymer solution
48
49 has a relatively high De. The fluid dynamics model described by Sinha-Ray et al. combines a
50
51
52 series of mass and momentum balances on the straight and perturbed or "whipping" components
53
54 of the jet to produce a time-dependent model of SBS jet dynamics. By accounting for the 3D
55
56 position of the perturbed jet in turbulent flowing air, as well as concentration and viscosity
57
58
59
60
ACS Paragon Plus Environment
7
ACS Applied Materials & Interfaces Page 8 of 55

1
2
3
changes due to solvent evaporation, this model is able to roughly estimate fiber size distribution
4
5
6 and approximate the positioning of fibers deposited on a moving collector. The model
7
8 moderately overestimates the fiber diameter distribution when compared to experimental data.
9
10
11
Further investigating the role of local turbulence in fiber formation, Lou et al. developed a 2D
12
13 computational fluid dynamics model to show the correlation of reliable fiber morphology with
14
15 low turbulent intensity of the gas velocity field.25 At gas pressures of 20 psi, a number
16
17
18 approximating what is used in many SBS devices, peak turbulent intensity of the simulated flow
19
20 field reached 35%. Experiments confirmed that at higher turbulent intensities polymer jets had
21
22 shorter straight segments and fibers showed a wider diameter distribution. Their findings
23
24
25 corroborate the frayed fiber morphology that has been observed by other groups at very high gas
26
27 pressures (58 psi).26 An optimal gas pressure for consistent fiber formation is likely one that
28
29
produces a flow field with velocities near the beginning of the turbulent regime.
30
31
32
33 These theoretical approaches indicate some of the critical parameters governing fiber
34
35 formation by SBS. Overlap concentration and polymer relaxation time are key factors in polymer
36
37
38 jet formation, implicating polymer concentration and molecular weight as key parameters
39
40 governing process design. The polymer jet dynamics model indicates that polymer concentration,
41
42 polymer flow rate, air flow rate, and nozzle dimensions will affect fiber diameter. The theoretical
43
44
45 importance of variables such as polymer concentration and air flow rate justify the empirical
46
47 approaches used by many groups to correlate critical parameters with fiber mat morphologies.
48
49
These empirical approaches will be reviewed in the next section.
50
51
52
53 2.2. Empirical Studies. 2.2.1. Polymer Concentration and Molecular Weight. Multiple studies
54
55 have examined the effects of polymer molecular weight or polymer concentration on fiber
56
57
58
formation and morphology. Confirming the critical role that overlap concentration Equation 1
59
60
ACS Paragon Plus Environment
8
Page 9 of 55 ACS Applied Materials & Interfaces

1
2
3
plays in enabling fiber formation, experiments have confirmed that SBS forms fibers above c*
4
5
6 (c>c*), “beads-on-a-string” near c* (c~c*), and corpuscular morphologies below c* (c<c*).19
7
8 Both poly(methyl methacrylate) molecular weight and concentration were varied across a range
9
10
11
spanning the estimated overlap concentration (using Equation 1 and parameters from the
12
13 literature) to visualize morphological changes under scanning electron microscopy (SEM)
14
15 (Figure 3B). Increasing concentration also produced a beads-on-a-string to fiber morphological
16
17
18 transition with poly(lactic-co-glycolic acid) (PLGA).5
19
20
21 Furthermore, within the domain c>c*, empirical studies show that fiber diameter increases
22
23 with increasing concentration. For poly(lactic acid) (PLA) fibers produced via SBS, increasing
24
25
26 polymer concentration from 4% w/v to 8% w/v produced fiber distributions with a greater
27
28 average diameter.26 This trend was confirmed with poly(ethylene oxide) (PEO) in a separate
29
30
31
study.25 However, increasing gas pressure reduced the effect of increased concentration.26 The
32
33 effects of gas pressure will be covered in depth in the next section.
34
35
36 2.2.2. Process Variables. Fiber diameter is only affected by polymer solution flow rate at low
37
38
39 gas pressures.26 However, low and high feed rates may cause jet instability and nozzle clogging,
40
41 respectively. Useable feed rates vary based on SBS device but generally range from 0.02 to 1 mL
42
43 min −1.2,4,26 Observations of nozzle clogging and jet instability at suboptimal operating conditions
44
45
46 have been reported by multiple research groups.2,5,26,27 Decreasing nozzle diameter decreases
47
48 fiber diameter.25 Increasing gas pressure produces a narrower fiber diameter distribution with
49
50
less variance and consistent fiber morphology, which is preferred for applications that require
51
52
53 precise fiber diameter (Figure 4A–C).5,26,28 However, increasing gas flow rate beyond the
54
55 optimal range may also cause a temperature decrease at the SBS device due to gas expansion,
56
57
58
which decreases temperature proportional to the volumetric expansion of the carrier gas.5 This
59
60
ACS Paragon Plus Environment
9
ACS Applied Materials & Interfaces Page 10 of 55

1
2
3
may cause poor solvent evaporation and fiber welding (Figure 4C). The observation of a stable
4
5
6 polymer jet has led researchers to search for an optimal set of polymer flow rates and air
7
8 pressures. Most blow spinning devices operate with polymer streams at room temperature,
9
10
11
however, temperature of the polymer stream may be increased to reduce the viscosity of the
12
13 polymer solution or increase solubility of the polymer.29
14
15
16 2.3. Choice of Polymer-Solvent System. Using information from the previous section, we
17
18
19 propose a set of general rules to guide polymer and solvent selection for SBS of polymer fibers.
20
21 The chosen solvent must be a good solvent for the polymer to be dissolved up to at least c*. The
22
23 concentration of the polymer in solution must be greater than c* to enable fiber formation, and
24
25
26 potentially higher to completely eliminate bead formation. Lastly, the polymer molecular weight
27
28 must be high enough to create entanglements between polymer chains, producing a sufficiently
29
30
31
high polymer relaxation time. Satisfying these minimum criteria will ensure the production of
32
33 polymer fibers. When using a solvent with exceptionally high or low inherent surface tension,
34
35 such as water, it may also be necessary to consider the contribution of surface tension to the
36
37
38 capillary forces that oppose fiber formation. It may be viable to increase solvent evaporation by
39
40 controlling the temperature and humidity of the surrounding environment during SBS, or using a
41
42 large working distance (50 cm).25,29 Solvent quality and evaporation rate may also affect the
43
44
45 crystallinity of the fibers formed by SBS.30 Filler content, such as nanoparticles, can increase the
46
47 viscosity of a polymer solution, and therefore will affect the morphology of fibers as described in
48
49
the previous sections.31 This has also been demonstrated in previous studies of electrospun
50
51
52 composite fibers, which showed that fillers can be used to modulate the viscosity of polymer
53
54 solutions, but fiber formation is still controlled by meeting the minimum polymer concentration
55
56
57
requirement of c*.32
58
59
60
ACS Paragon Plus Environment
10
Page 11 of 55 ACS Applied Materials & Interfaces

1
2
3
3. STRUCTURE AND PROPERTIES OF FIBERS
4
5
6
7 3.1. Morphology. Controlling the microstructure, morphology, and alignment of fibers enables
8
9 the fabrication of multifunctional fibrous materials, including fibrous composites, tissue
10
11
12 engineering scaffolds with enhanced regenerative potential, and nonwoven textiles with
13
14 increased surface area and desirable transport properties.33–36 To address these needs, SBS has
15
16 been used to easily create fibers conferred with unique morphologies, various microstructural
17
18
19 features, and diverse mechanical properties. Fibers with diameters ranging from ~100 nm to
20
21 greater than 1 µm have been fabricated by SBS. Aligned fiber mats can be created by spinning
22
23 onto a rolling collector, similar to meltspinning and electrospinning.3 By modulating the
24
25
26 concentration of polymer in the polymer solution, SBS can be used to produce polymer
27
28 constructs with fibrous, beads on a string, and corpuscular morphologies for enhanced surface
29
30
31
properties, such as increased omniphobicity.19 Porosity and fiber branching can be controlled
32
33 using process variables such as polymer concentration and polymer blending.26,37
34
35
36 SBS and electrospinning produce fiber mats with different morphological characteristics.
37
38
39 Overall porosity (77–95%) and pore size (8–17 µm) of SBS scaffolds are greater than those of
40
41 scaffolds produced by electrospinning with similar polymers (67% and 3 µm, respectively).2
42
43 Tutak et al. demonstrated that SBS may produce fiber bundles, a non-uniform morphology not
44
45
46 associated with electrospinning (Figure 5A–B). This observation was confirmed by Bolbasov et
47
48 al. using polyvinylidene fluoride-trifluroethylene copolymer (VDF-TeFE).38 Tutak et al. also
49
50
observed that SBS produces fibers with a tighter diameter distribution than electrospinning, an
51
52
53 observation that was confirmed by Oliveira et al.39 Locally non-uniform structural control may
54
55 provide a means to investigate cell response to specific three-dimensional biomimetic structures,
56
57
58
such as aligned fibrils, which have been shown to modulate cell migration and protein expression
59
60
ACS Paragon Plus Environment
11
ACS Applied Materials & Interfaces Page 12 of 55

1
2
3
in collagen matrices.40 To further increase scaffold porosity and enhance cell infiltration,
4
5
6 Medeiros et al. recently developed cryogenic SBS.41 This technique simultaneously incorporates
7
8 ice spheres into freeze-dried fibers, creating macroporous fibers when the scaffold is deployed
9
10
11
and the ice melts.
12
13
14 3.2. Mechanical Properties. The mechanical properties of a fiber scaffold can be critical to its
15
16 end use, e.g. as a nonwoven textile, or a feature used to increase effectiveness, such as in cell
17
18
19 infiltration. Young's modulus, failure strain, and failure stress, derived from a stress-strain curve,
20
21 are frequently reported for materials of interest. However, few studies have comprehensively
22
23 examined the mechanical properties of SBS fiber scaffolds. One study has compared the
24
25
26 mechanical properties of polycaprolactone (PCL) scaffolds produced by SBS and
27
28 electrospinning, and found that electrospinning produced fiber mats with approximately 10 times
29
30
31
higher Young's modulus (Figure 5C–D).2 SEMs showed greater fiber entanglement in the
32
33 electrospun scaffolds, which likely caused them to have greater stiffness (Figure 5A–B). Another
34
35 compared the mechanical properties of VDF-TeFE copolymers fabricated by SBS and
36
37
38 electrospinning.38 SBS again produced mats with local fiber bundles, unlike electrospinning, and
39
40 lower tensile strength. Electrospinning processes can also be tuned to align and organize fibers
41
42 using additional equipment, such as rotating collectors or counter electrodes, allowing an
43
44
45 electrospinning setup to produce similar local orientation or levels of entanglement.42–45 Block
46
47 copolymer fiber mats fabricated by SBS have demonstrated notable elasticity and strain recovery
48
49
(Figure 5E).1 Energy dissipation studies showed the modes of failure for fiber mats produced by
50
51
52 SBS, which are fiber rearrangement, broken fiber connections, and individual fiber deformation.1
53
54 Further investigation is required into the connection between the bundled morphology of SBS
55
56
57
fiber mats and their mechanical properties. Fiber bundles may be related to turbulent air flow
58
59
60
ACS Paragon Plus Environment
12
Page 13 of 55 ACS Applied Materials & Interfaces

1
2
3
around the polymer jet and subsequent bending instability, which causes multiple streams of
4
5
6 solution to emerge from a single nozzle (Figure 6A). Multiple fibers may be formed
7
8 simultaneously and thus deposited in the same area with alignment. Lower gas pressures may
9
10
11
reduce these instabilities (Figure 6B–C).46 Fiber mats with lower modulus may better
12
13 approximate the mechanical properties of soft biological materials such as fibrin (1–10 MPa) or
14
15 human skin (0.1–1 MPa).47,48
16
17
18
19 3.3. Multicomponent Polymer Fibers. Multicomponent polymer mixtures and coaxial SBS
20
21 setups with concentric nozzles have been used to create polymer fibers with well-defined
22
23 heterogeneous polymer distributions. Oliveira et al. used a blend of PEO and PLA to fabricate
24
25
26 fibers with a core of amorphous PLA and shell of semi-crystalline PEO when blended at a 1:1
27
28 ratio (Figure 7A).49 Core-shell amorphous polymer fibers have been created from coaxial SBS of
29
30
31
poly(methyl methacrylate) (PMMA) and polyacrylonitrile (PAN), soy protein and nylon-6, and
32
33 wood cellulose pulp and PEO.50–52 In the last two examples, the polymer shell was employed to
34
35 stabilize the formation of the fiber core, which could not be formed alone. Core-shell fibers with
36
37
38 epoxy precursors or other self-healing monomers loaded in the core exhibited self-healing
39
40 properties and increased fatigue strength.53–55
41
42
43 Polymer blends have also been used with SBS to tailor the properties of homogenous polymer
44
45
46 fibers. Liu et al. blended chitosan and poly(vinyl alcohol) (PVA) with varying amounts of
47
48 crosslinker to create swellable hydrogel nanofibers (Figure 7B–C).56 Behrens et al. blended
49
50
PLGA and poly(ethylene glycol) (PEG) in various compositions to tune the glass transition
51
52
53 temperature of polymer fibers, though this also affected the fiber diameter and morphology
54
55 (Figure 7D–F).57 PEG content also modulates swelling and permeability in chitosan/PLA blend
56
57
58
fibers.58 Polymer blending using polymers of different molecular weights can alter degradation
59
60
ACS Paragon Plus Environment
13
ACS Applied Materials & Interfaces Page 14 of 55

1
2
3
rate.37 Mixtures of polymer with non-polymer additives such as zirconium-modified amorphous
4
5
6 calcium phosphate and low surface energy cage molecule 1H,1H,2H,2H-heptadecafluorodecyl
7
8 polyhedral oligomeric silsesquioxane have been used to tailor surface properties of fibers made
9
10
11
by SBS.19,59 Research groups have also used SBS to airbrush fibers with liquid crystal cores and
12
13 polymer shells formed through spontaneous phase separation during the spraying process.60
14
15
16 4. SBS APPLICATIONS AND INNOVATIONS
17
18
19
20 SBS allows for portable, conformal fiber deposition on any substrate. This dramatically expands
21
22 the number of possible applications for nanofiber-based technologies. Previously, possible
23
24
targets for fiber deposition were limited by restraints inherent to the technique: electrospinning
25
26
27 requires an applied voltage and conductive target, while industrial processes require cumbersome
28
29 equipment. As a result, the most impactful research using SBS has studied direct deposition onto
30
31
32
the target of interest, especially on biological substrates. The following discussion will highlight
33
34 these applications, how each application has improved upon or integrated conventional
35
36 techniques in its field, and the broader significance and future impact that we project each
37
38
39 application will have.
40
41
42 4.1. Biomedical Applications of Polymer Fibers. Polymer fiber mats produced by SBS are
43
44 porous, making them ideal for use as a cell scaffold. Fibrous scaffolds are designed to foster cell
45
46
47 proliferation, differentiation, and infiltration. SBS can produce nanofiber scaffolds capable of
48
49 culturing human bone marrow stromal cells (hBMSCs)2. The porosity and pore size of SBS
50
51
52
scaffolds facilitate cell infiltration. A direct comparison of hBMSCs cultured on blow spun and
53
54 electrospun scaffolds showed that hBMSCs cultured on blow spun scaffolds penetrated 78.75 ±
55
56
18.46 µm, significantly deeper than on electrospun scaffolds (34.75 ± 8.77 µm).59 Mesenchymal
57
58
59
60
ACS Paragon Plus Environment
14
Page 15 of 55 ACS Applied Materials & Interfaces

1
2
3
stem cell (MSC) proliferation and differentiation was also measured on VDF-TeFE scaffolds
4
5
6 produced by SBS and electrospinning.38 MSCs differentiated into a higher fraction of
7
8 proliferative phenotype when cultured on SBS scaffolds and produced more total cells (Figure
9
10
11
8A–C). Polymer nanofiber scaffolds were shown to increase cellular peroxidase activity and
12
13 influence organelle positioning versus polymer films of the same composition.61 SBS has also
14
15 been used to fabricate PLA implants loaded with dibasic calcium phosphate dehydrate filler.62
16
17
18 These implants were shown to be oncologically safe as they do not stimulate the growth of
19
20 tumors adjacent to the implant.
21
22
23 Tissue engineering and regenerative medicine will benefit from the versatility of conformal
24
25
26 deposition by SBS. Target wounds and defects for tissue replacement have varying size and
27
28 geometry, making adaptable materials fabrication a necessity for the clinical future of these
29
30
31
fields. Furthermore, as a field, tissue engineering has devoted significant resources to developing
32
33 custom materials fabrications systems, such as 3D printers and injectable hydrogels.63 SBS can
34
35 meet a subset of these needs by providing timely, on-demand polymer fiber deposition.5 SBS is
36
37
38 also compatible with a variety of additives that have been shown to enhance regenerative
39
40 potential, such as osteogenic zirconium-modified amorphous calcium phosphate.59 Nanoparticles
41
42 of bioactive glass have also been used as additive in SBS scaffolds and delivered using a burst
43
44
45 release.41 Solvent toxicity is a problem with electrospun fiber scaffolds that has been addressed
46
47 by melt-electrospinning processes, which spin polymer melts to avoid the need for solvent.64
48
49
SBS inherently solves this problem by volatilizing the spinning solvent: Acetone showed no
50
51
52 cytotoxicity when sprayed at a distance of 10 cm directly onto cell culture plates, both during
53
54 fiber deposition and alone.5 SBS can be performed aseptically in a sterile environment.65
55
56
57
58
59
60
ACS Paragon Plus Environment
15
ACS Applied Materials & Interfaces Page 16 of 55

1
2
3
Medical applications are uniquely suited for direct fiber deposition in targets and will also be
4
5
6 able to utilize the properties of SBS that make it particularly biocompatible: low toxicity, high
7
8 porosity, and compatibility with biodegradable materials. SBS has been used to directly coat
9
10
11
medical devices with lubricant-infused poly(styrene-b-isobutylene-b-styrene) (SIBS)
12
13 microfibers, which can resist thrombosis and bacterial attachment with low cytotoxicity.66 The
14
15 addition of silicon oil prevented the adhesion of blood cells (Figure 8D–E). Some promising
16
17
18 biodegradable and absorbable polymers previously used in US Food and Drug Administration-
19
20 approved products have been used with SBS (Table 1). An initial investigation of PLGA fibers
21
22 for biomedical applications demonstrated nontoxicity of the SBS process, hydrolytic degradation
23
24
25 of the fibers leading to morphological changes, interactions between nanofibers and blood, and
26
27 in situ deposition in porcine models of lung resection, intestinal anastomosis, liver injury, and
28
29
diaphragmatic hernia.5 In situ conformal deposition by SBS from a commercial airbrush allowed
30
31
32 for complete coverage of tissues with polymer fibers (Figure 8F), which transitioned to an
33
34 adhesive film for use as a surgical sealant (Figure 8G).57 Polymer fibers have also been created
35
36
37 with SBS for antibacterial activity using biopolymers, PLA/polyvinylpyrrolidone (PVP) blends,
38
39 and chitosan/PLA/PEG blends loaded with antibacterial additives.56,67,68 Controlled drug delivery
40
41 has been achieved using nanostructured membranes and core-shell fibers.69,70 Desorption was
42
43
44 determined to be the limiting factor in release of a hydrophobic drug, and poragens such as PEG
45
46 were used to control release rate.71 An electrochemical glucose biosensor and metal ion sensor
47
48 for potable water were also developed based on a platform of fibers spun from a mixture of PLA
49
50
51 and carbon.72,73
52
53
54 4.2. Polymer Fibers for Coatings, Textiles, and Electronics. SBS has its conceptual roots in
55
56
57
fiber spinning techniques that are commonly used in industrial fiber production processes, such
58
59
60
ACS Paragon Plus Environment
16
Page 17 of 55 ACS Applied Materials & Interfaces

1
2
3
as polymer solution spinning from a spinneret to create textiles74. Many of the same physical
4
5
6 principles govern both processes. Due to its simplicity and ability to directly deliver conformal
7
8 fibers, it has been investigated for creating functional polymer nanofiber coatings, nonwoven
9
10
11
textiles, and stretchable electronics. These investigations have shown that SBS can produce
12
13 fibrous materials with high precision, efficiency, and reliability.
14
15
16 Fibrous coatings may utilize porous microstructures with re-entrant surface features to
17
18
19 maximize omniphobicity.75 SBS can produce superomniphobic fibrous coatings with finely-
20
21 tuned corpuscular microstructure controlled by polymer concentration in solution, with the
22
23 eventual possibility of direct, conformal delivery.19 It proved to be an efficient method for
24
25
26 nanofiber generation, safely and cost-effectively delivering nanofibers at high deposition rates
27
28 with the potential for further process scaling.2,4 SBS has also been used for filter and membrane
29
30
31
applications. Shi et al. prepared air filters from nylon-6 and Lee et al. prepared water purification
32
33 membranes from nylon-6 fibers entangled with graphene flakes during the spraying process.76,77
34
35 PAN-based activated carbon fibers were fabricated with SBS and used for CO2 adsorption and
36
37
38 phenol adsorption.78,79 Ultrafine 20–50 nm fibers produced by SBS can be used for nanoparticle
39
40 filtration.80 Poly(ether sulfone)/Nafion and poly(ether ether ketone)/Nafion composite proton
41
42 exchange membranes were also fabricated using SBS fibers impregnated with Nafion solution.81–
43
44 83
45
46
47
48 SBS has been used to fabricate stretchable conductive materials, via silver nanoparticle-based
49
50
solution-processing techniques.1 By conformally depositing this material on a hand and
51
52
53 measuring electronic resistance (Figure 9A), this research shows the potential of using SBS to
54
55 develop wearable smart materials. Fibrous composite electronic devices show enhanced
56
57
58
stretchability and conductivity. SBS enables their deposition on nonplanar substrates, which has
59
60
ACS Paragon Plus Environment
17
ACS Applied Materials & Interfaces Page 18 of 55

1
2
3
the potential to enhance commercial applicability. SBS can also serve as the framework for
4
5
6 advanced processing methods, such as carbon nanofiber production (Figure 9B–C).51,84
7
8 Similarly, yarns of carbon nanofibers were produced using PAN as a precursor polymer.85 By
9
10
11
spinning onto two rods, a yarn was formed. SBS fiber mats have also been used as a precursor
12
13 for alumina, zirconia, and mullite fiber preparation.86–88 Y-Ba-Cu-O oxide ceramic fibers were
14
15 fabricated for use in superconducting applications.89 Ceramic nanofibers of TiO2 and ZnO with
16
17
18 high surface area have been fabricated using SBS from a mixture of Ti or Zn sources and
19
20 polymer in solution, followed by post-processing.90 The primary advantage of using SBS to
21
22 fabricate these materials is increased production rate, which will make processes more cost-
23
24
25 efficient and thus more accessible to researchers and markets. Depending on the type of ceramic
26
27 fiber, production rates typically increase by a factor of 5–10 when using SBS compared to
28
29
conventional electrospinning.89,91
30
31
32
33 5. FUTURE DIRECTIONS
34
35
36 SBS provides opportunities for expanding upon the foundation of nanofiber-based research
37
38
39 investigated in the literature using the electrospinning technique. SBS has fewer constraints than
40
41 electrospinning. Because fiber formation is not driven by an electric field, there is no effect on
42
43 fiber diameter from the electrical conductivity of the polymer solution, and subsequently, no
44
45
46 incentive to use highly toxic fluorinated solvents. Further, the simple apparatus required for SBS
47
48 will allow researchers to investigate new applications for nanofibrous and microfbrous materials.
49
50
Because they can be deposited on demand, SBS can feasibly create conformal coatings, textiles,
51
52
53 and materials with custom geometries from a scalable self-contained process. This is a necessary
54
55 innovation for the advancement of fields such as wearable electronics and healthcare
56
57
58
materials.92,93
59
60
ACS Paragon Plus Environment
18
Page 19 of 55 ACS Applied Materials & Interfaces

1
2
3
Direct deposition onto living targets requires further investigation as a method for creating
4
5
6 biological scaffolds, especially given the differences in fiber morphology observed between
7
8 electrospinning and SBS. Previous research has demonstrated the viability of cells deposited
9
10
11
from an airbrush.94,95 Crucial parameters such as polymer concentration, solution viscosity, and
12
13 spraying technique may affect cell viability but do not prevent the simultaneous deposition of
14
15 cells and material.96 Biological materials, such as collagen and silk, merit further investigation as
16
17
18 a material source for SBS given their outstanding mechanical properties, particularly if they can
19
20 be spun in their native folded state and then used to fabricate low cost biomimetic materials.97 A
21
22 number of biological materials have been spun into fibers by SBS, including soy protein, fish
23
24
25 sarcoplasmic protein, and cellulose acetate (Table 2).52,98,99 In conjunction, novel modifications
26
27 to the general SBS technique can allow for spinning from new solvents, such as water or
28
29
dimethyl carbonate, and the ability to deposit cells (Figure 10A–B).29,65,100
30
31
32
33 SBS will enable researchers to develop materials that are translatable to commercial markets.
34
35 Scalability and safety are necessary for the widespread applicability of nanofibrous materials.9
36
37
38 Beyond simple advantages in portability due to the fact that SBS can deposit fibers onto any
39
40 target, SBS is a more cost-effective and rapid technique for generating nanofibers than
41
42 electrospinning2. Costs are minimized when a gravity-fed or siphon-fed airbrush is used, while
43
44
45 deposition rates are on the order of 10 times faster than electrospinning. Various research groups
46
47 have increased process scale by using multiple nozzles27,101, and grids65 (Figure 10C).
48
49
Mahalingam et al. complexed SBS with centrifugal spinning to increase production102. Safety
50
51
52 concerns related to the usage of toxic solvents can also be minimized by using volatile solvents
53
54 with limited toxicity that evaporate before accumulating at the surface of the fiber collector. SBS
55
56
57
is also adaptable to spinning many different types of polymer-composite and polymer-composite
58
59
60
ACS Paragon Plus Environment
19
ACS Applied Materials & Interfaces Page 20 of 55

1
2
3
precursor mixtures (Table 3). This enables the cost-effective fabrication of highly functional
4
5
6 nonwoven nanotextiles, which may have increased durability or provide optical responses to
7
8 external stimuli.103
9
10
11
12 Further research into the fundamental parameters governing SBS fiber formation is required. It
13
14 is clear that a multitude of factors such as polymer concentration and gas pressure influence fiber
15
16 diameter and morphology, but the effects of solvent evaporation rate, polymer molecular weight,
17
18
19 polymer blends and solution additives, and nozzle type require further investigation. It is also
20
21 important to clarify the differences between variations of the SBS technique.104 Research on
22
23 well-studied fiber spinning techniques similar to SBS may provide insight on how to answer
24
25
26 these questions.9
27
28
29 6. CONCLUSION
30
31
32
33
SBS is a fabrication technique that enables on-demand, in situ fiber generation using a simple
34
35 apparatus. Theoretical and empirical models of fiber production using this technique have
36
37 revealed basic relationships between polymer solution concentration, gas pressure, and fiber
38
39
40 diameter. These studies also emphasize the importance of overlap concentration and molecular
41
42 weight in determining the spinnability of a polymer solution. SBS materials have a bundled
43
44 morphology and unique mechanical properties. Special techniques and multicomponent blends
45
46
47 can be used to produce functional architectures, such as core-shell fibers. A range of applications
48
49 involving polymer nanofibers generated by SBS in biomaterials, tissue engineering, textiles, and
50
51
52
composites have been highlighted. In the future, SBS will be most valuable to the research
53
54 community as an avenue for exploring new combinations of polymers and solvents not
55
56
57
58
59
60
ACS Paragon Plus Environment
20
Page 21 of 55 ACS Applied Materials & Interfaces

1
2
3
accessible to electrospinning, and as a technique for conveniently producing translatable fibrous
4
5
6 materials.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
21
ACS Applied Materials & Interfaces Page 22 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 1. A) Schematic for a solution blow spinning device showing concentric nozzles.
42
43
44
Adapted with permission.1 Copyright 2015, American Chemical Society. B) A general solution
45
46 blow spinning process diagram with rolling collector. Adapted with permission.3 Copyright
47
48 2009, Wiley-VCH. C) Image of direct deposition of poly(styrene-block-isoprene-block-styrene)
49
50
51 block copolymer fibers using a homemade solution blow spinning device. The homemade device
52
53 was made from a transfer pipet and a flat tipped 18G needle (inset). Adapted with permission.1
54
55 Copyright 2015, American Chemical Society. D) A commercial airbrush used for solution blow
56
57
58 spinning. Adapted with permission.2 Copyright 2013, Elsevier.
59
60
ACS Paragon Plus Environment
22
Page 23 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 2. The number of publications on solution blow spinning and related topics has increased
19
20
since its widespread exposure began in 2009. Data for 2016 is current as of September.
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
23
ACS Applied Materials & Interfaces Page 24 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Figure 3. A) Plot indicating morphology of poly(methyl methacrylate) (PMMA) sprayed using a
39
40
41 solution blow spinning (SBS) apparatus at various concentrations and molecular weights. The
42
43 estimated overlap concentration (c*) is indicated by the dashed line. Scanning electron
44
45
46 microscopy (SEM) image of PMMA fibers formed at high molecular weight but below overlap
47
48 concentration. Scale bar represents 50 µm. B) SEM images of 50/50 wt.%
49
50 PMMA/1H,1H,2H,2H-heptadecafluorodecyl polyhedral oligomeric silsesquioxane (PMMA: Mw
51
52
53 = 593 kDa, PDI = 2.69) blends sprayed using an SBS apparatus at increasing concentrations of
54
55 PMMA in solution. Scale bars represent 50 µm, 100 µm, and 50 µm, respectively. Adapted with
56
57 permission.19 Copyright 2011, Elsevier.
58
59
60
ACS Paragon Plus Environment
24
Page 25 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 Figure 4. A–C) Scanning electron microscopy (SEM) images of fiber morphologies produced by
16
17
18
solution blow spinning 10% w/v poly(lactic-co-glycolic acid) in acetone at CO2 flow rates of 10
19
20 SCFH (A), 13 SCFH (B), and 15 SCFH (C). Scale bars correspond to 20 µm. Adapted with
21
22 permission.5 Copyright 2014, American Chemical Society.
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
25
ACS Applied Materials & Interfaces Page 26 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 5. Morphology, Young’s modulus, and cyclic testing of solution blow spinning (SBS)
34
35
36 fibers. A) Fiber bundles produced by SBS of a 4 wt.% polymer solution of poly(caprolactone)
37
38 (PCL) (80 kDa) in chloroform. B) Fibers produced by electrospinning a 4 wt.% polymer solution
39
40
41 of PCL (80 kDa) in 3:1 chloroform:methanol by mass. A−B adapted with permission.2 Copyright
42
43 2013, Elsevier. C) Comparison of stress-strain curves for PCL (80000 kDa) fiber mats fabricated
44
45 using electrospinning and SBS. D) Young’s modulus for fiber mats produced using SBS and
46
47
48 electrospinning PCL fibers. C−D adapted with permission.2 Copyright 2013, Elsevier. E)
49
50 Stress/strain cycling curves at varying maximum strain values for poly(styrene-block-isoprene-
51
52
53 block-styrene) fiber mats fabricated by SBS. Adapted with permission.1 Copyright 2015,
54
55 American Chemical Society.
56
57
58
59
60
ACS Paragon Plus Environment
26
Page 27 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 6. Images showing the effects of turbulence on polymer jets used for solution blow
31
32
33 spinning (SBS). A) High-speed photograph of a polymer solution cone forming at the tip of a
34
35 SBS device. Adapted with permission.3 Copyright 2009, Wiley-VCH. B−C) High-speed
36
37
38 photograph comparing the length of the straight region (in red box) of the polymer jet produced
39
40 by SBS at gas pressures of 1.121 atm (B) and 1.363 atm (C). Adapted with permission.46
41
42 Copyright 2014, American Chemical Society.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
27
ACS Applied Materials & Interfaces Page 28 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 7. Various multicomponent fibers produced by solution blow spinning (SBS). A)
35
36
37 Transmission electron microscopy images of core-shell fibers produced by SBS from a 1:1 blend
38
39 of poly(lactic acid) and poly(ethylene glycol) (PEG) at 6% w/v in solution. Adapted with
40
41 permission.49 Copyright 2013, Wiley-VCH. B–C) Scanning electron microscopy (SEM) images
42
43
44 of fibers produced by SBS from a 1:1 blend of chitosan and poly(vinyl alcohol) at 8% w/v in
45
46 solution with 7% ethylene glycol diglycidyl ether crosslinker (relative to mass of polymer). SEM
47
48
images were taken after drying in a vacuum oven for 3 hours (B), and after drying then swelling
49
50
51 in saline solution for 8 hours (C). Adapted with permission.56 Copyright 2014, Elsevier. D)
52
53 Fibers produced by SBS from a 10% w/v poly(lactic-co-glycolic acid) (PLGA) in acetone
54
55
56 solution. E) Fibers produced by SBS from a 10% w/v PLGA, 5% w/v PEG blend in solution. F)
57
58
59
60
ACS Paragon Plus Environment
28
Page 29 of 55 ACS Applied Materials & Interfaces

1
2
3
Fiber diameter produced by SBS from a 10% w/v PLGA solution increases with increasing PEG
4
5
6 blend content. Adapted with permission.57 Copyright 2015, Wiley-VCH.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
29
ACS Applied Materials & Interfaces Page 30 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Figure 8. Biomaterials applications of solution blow spinning (SBS). A–C) Fluorescent
33
34 microscopy images showing MSCs cultured on a flat cell culture control surface (A),
35
36 polyvinylidene fluoride-trifluroethylene copolymer (VDF-TeFE) fibers produced by
37
38
39 electrospinning (B), and VDF-TeFE fibers produced by SBS (C). Scale bars represent 100 µm.
40
41 Adapted with permission.38 Copyright 2016, Elsevier. D–E) Field emission scanning electron
42
43
microscopy (FESEM) images showing blood cell and platelet attachment to poly(styrene-b-
44
45
46 isobutylene-b-styrene) (SIBS) fibers (D) and SIBS fibers infused with silicon oil (E) after
47
48 incubation with whole blood. Scale bars represent 10 µm. Adapted with permission.66 Copyright
49
50
51 2016, American Chemical Society. F) Intestinal anastomosis being coated in fibers produced by
52
53 SBS from a poly(lactic-co-glycolic acid) (PLGA) and poly(ethylene glycol) (PEG) blend. G)
54
55 Intestinal anastomosis 2 minutes after being coated showing the thermal transition of PLGA-
56
57
58 PEG fibers. F–G adapted with permission.57 Copyright 2015, Wiley-VCH.
59
60
ACS Paragon Plus Environment
30
Page 31 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 9. Composite and post-processing techniques enabled by solution blow spinning (SBS).
19
20 A) Glove coated in poly(styrene-block-isoprene-block-styrene) elastomeric fibers patterned with
21
22
23 conductive bands of silver nanoparticles showing resistance measured across each line
24
25 corresponding to hand gestures. Adapted with permission.1 Copyright 2015, American Chemical
26
27 Society. B–C) Side (B) and cross section (C) of carbon tubes processed from core-shell
28
29
30 poly(methyl methacrylate)-poly(acrylonitrile) fibers produced by SBS. Adapted with
31
32 permission.51 Copyright 2010, Elsevier.
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
31
ACS Applied Materials & Interfaces Page 32 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 10. Future directions for solution blow spinning (SBS) research. A–B) Images of
22
23 nanofibers produced by SBS from a 15% polyvinylpyrrolidone solution in fetal bovine serum
24
25
26 sprayed without (A) and with (B) mouse renal epithelial cells in the solution. Adapted with
27
28 permission.65 Copyright 2012, RSC Publishing. C) Multinozzle SBS process using a die with 41
29
30
nozzles per row and 8 rows. Adapted with permission.27 Copyright 2016, American Chemical
31
32
33 Society.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
32
Page 33 of 55 ACS Applied Materials & Interfaces

1
2
3
Table 1. Biodegradable polymers and solvent combinations used for solution blow spinning.
4
5
6
7 Polymer Solvent Concentration Gas Pressure Reference
8 (% w/v) (psi)
9
10 PCL Chloroform 8 35 2
11
12 PLA Chloroform/acetone, 6 35, 29–58 2,26
13 Chloroform/methanol
14
15 PLGA Acetone 10 N/A 5
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
33
ACS Applied Materials & Interfaces Page 34 of 55

1
2
3
Table 2. Biopolymer and solvent combinations used for solution blow spinning.
4
5
6
7 Polymer Solvent Concentration Gas Pressure Reference
8 (wt %) (psi)
9
10 Bovine serum Water 10/10 29 98
11 albumin/Poly(vinyl
12 alcohol)
13
14 Cellulose LiCl/ 2 36 50
15
16
Dimethylacetamide
17 (DMAc)
18
19 Cellulose/Poly(ethylene LiCl/DMAc 2/10 36 50
20 oxide)
21
22 Cellulose N,N- 5/5 29 98
23 acetate/Polyacetonitrile dimethylacetamide
24 (DMF)
25
26 Chitosan/Poly(vinyl Formic acid 3.2/3.2 72 56
27
28 alcohol)
29
30 Fish sarcoplasmic Formic acid 1.75–17.5/ 60 99
31 protein/Nylon-6 1.75–17.5
32
33 Lignin Formic acid 5–14 29 98
34
35 Hydroxypropyl Trifluoroethanol 9 80 67
36 methylcellulose
37 /Poly(lactic acid)
38
39 Silk sericin/Nylon-6 Formic acid 12/12 29 98
40
41 Soy protein/Nylon-6 Formic acid 9/14 N/A 52
42
43 Zein/Nylon-6 Formic acid 9–22/11–13 29 98
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
34
Page 35 of 55 ACS Applied Materials & Interfaces

1
2
3
Table 3. Polymer-composite mixtures used for solution blow spinning.
4
5
6
7 Polymer-composite Solvent Concentration Gas Pressure Reference
8 mixture (wt %) (psi)
9
10 AlCl3∙6H2O/Silica Water 15/2/.4 1–10 86
11 particles/PVA
12
13 Al(NO3)∙9H2O/Si(OC2H5)4 THF N/A/10 40 88
14
15
/PVC
16
17 PCL/Zirconium-modified Chloroform 4/0–20 30–40 59
18 amorphous calcium
19 phosphate (Zr-ACP)
20
21 PLA/Carbon nanotubes Chloroform/ 6/0–3 58 72
22 acetone
23
24 PLA/Nano-bioactive glass Dimethyl 3–13/4–8 20 41
25 carbonate
26
27
PLA/Zr-ACP Acetone 8/0–20 30–40 59
28
29
PMMA/Zr-ACP Acetone 10/0–20 30–40 59
30
31
32
PVP/ZrOCl3 Water 2–5/9–15 6–17 87
33
34
PVP/YBCO Methanol/acetic 4/11 19 89
35 acid/propionic
36 acid
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
35
ACS Applied Materials & Interfaces Page 36 of 55

1
2
3
ASSOCIATED CONTENT
4
5
6
7 AUTHOR INFORMATION
8
9
10 Corresponding Author
11
12 E-mail: kofinas@umd.edu
13
14
15
16 Author Contributions
17
18 The manuscript was written through contributions of all authors. All authors have given approval
19
20
21 to the final version of the manuscript.
22
23
24 ACKNOWLEDGMENT
25
26
27
Research reported in this publication was supported by the National Institute of Biomedical
28
29 Imaging and Bioengineering of the National Institutes of Health under Award No.
30
31 R01EB019963. The content is solely the responsibility of the authors and does not necessarily
32
33
34 represent the official views of the National Institutes of Health. The authors would like to thank
35
36 Dr. Omar Ayyub for helpful discussions.
37
38
39 REFERENCES
40
41
42
(1) Vural, M.; Behrens, A. M.; Ayyub, O. B.; Ayoub, J. J.; Kofinas, P. Sprayable Elastic
43 Conductors Based on Block Copolymer Silver Nanoparticle Composites. ACS Nano 2015,
44 9 (1), 336–344.
45 (2) Tutak, W.; Sarkar, S.; Lin-Gibson, S.; Farooque, T. M.; Jyotsnendu, G.; Wang, D.; Kohn,
46 J.; Bolikal, D.; Simon, C. G. The Support of Bone Marrow Stromal Cell Differentiation by
47
Airbrushed Nanofiber Scaffolds. Biomaterials 2013, 34 (10), 2389–2398.
48
49 (3) Medeiros, E. S.; Glenn, G. M.; Klamczynski, A. P.; Orts, W. J.; Mattoso, L. H. C. Solution
50 Blow Spinning: A New Method to Produce Micro- and Nanofibers from Polymer
51 Solutions. J. Appl. Polym. Sci. 2009, 113 (4), 2322–2330.
52 (4) Polat, Y.; Pampal, E. S.; Stojanovska, E.; Simsek, R.; Hassanin, A.; Kilic, A.; Demir, A.;
53 Yilmaz, S. Solution Blowing of Thermoplastic Polyurethane Nanofibers: A Facile Method
54
55
to Produce Flexible Porous Materials. J. Appl. Polym. Sci. 2016, 133 (9), 43025. DOI:
56 10.1002/APP.43025.
57
58
59
60
ACS Paragon Plus Environment
36
Page 37 of 55 ACS Applied Materials & Interfaces

1
2
3
(5) Behrens, A. M.; Casey, B. J.; Sikorski, M. J.; Wu, K. L.; Tutak, W.; Sandler, A. D.;
4
5 Kofinas, P. In Situ Deposition of PLGA Nanofibers via Solution Blow Spinning. ACS
6 Macro Lett. 2014, 3 (3), 249–254.
7 (6) Son, W. K.; Youk, J. H.; Lee, T. S.; Park, W. H. The Effects of Solution Properties and
8 Polyelectrolyte on Electrospinning of Ultrafine Poly(ethylene Oxide) Fibers. Polymer
9 2004, 45 (9), 2959–2966.
10
11
(7) Luo, C. J.; Nangrejo, M.; Edirisinghe, M. A Novel Method of Selecting Solvents for
12 Polymer Electrospinning. Polymer 2010, 51 (7), 1654–1662.
13 (8) Greiner, A.; Wendorff, J. H. Electrospinning: A Fascinating Method for the Preparation of
14 Ultrathin Fibers. Angew. Chem. Int. Ed. 2007, 46 (30), 5670–5703.
15 (9) Luo, C. J.; Stoyanov, S. D.; Stride, E.; Pelan, E.; Edirisinghe, M. Electrospinning versus
16
Fibre Production Methods: From Specifics to Technological Convergence. Chem. Soc.
17
18 Rev. 2012, 41 (13), 4708–4735.
19 (10) Badrossamay, M. R.; McIlwee, H. A.; Goss, J. A.; Parker, K. K. Nanofiber Assembly by
20 Rotary Jet-Spinning. Nano Lett. 2010, 10 (6), 2257–2261.
21 (11) Benavides, R. E.; Jana, S. C.; Reneker, D. H. Nanofibers from Scalable Gas Jet Process.
22 ACS Macro Lett. 2012, 1 (8), 1032–1036.
23
24
(12) Weitz, R. T.; Harnau, L.; Rauschenbach, S.; Burghard, M.; Kern, K. Polymer Nanofibers
25 via Nozzle-Free Centrifugal Spinning. Nano Lett. 2008, 8 (4), 1187–1191.
26 (13) Dean, A. R.; Emilio, R. J. Process and Apparatus for Flash Spinning of Fibrillated
27 Plexifilamentary Material. US3227794 A, January 4, 1966.
28 (14) Kim, S. Y.; Purnama, P.; Kim, S. H. Fabrication of Poly(l-Lactide) Fibers/sheets Using
29
Supercritical Fluid through Flash-Spinning Process. Macromol. Res. 2010, 18 (12), 1233–
30
31 1236.
32 (15) Stojanovska, E.; Canbay, E.; Pampal, E. S.; Calisir, M. D.; Agma, O.; Polat, Y.; Simsek,
33 R.; Gundogdu, N. A. S.; Akgul, Y.; Kilic, A. A Review on Non-Electro Nanofibre
34 Spinning Techniques. RSC Adv 2016, 6 (87), 83783–83801.
35 (16) Graessley, W. W. Polymer Chain Dimensions and the Dependence of Viscoelastic
36
37 Properties on Concentration, Molecular Weight and Solvent Power. Polymer 1980, 21 (3),
38 258–262.
39 (17) Paul J. Flory. Statistical Mechanics of Chain Molecules; Interscience Publishers, 1969.
40 (18) Paul J. Flory. Principles of Polymer Chemistry; Cornell University Press, 1953.
41 (19) Srinivasan, S.; Chhatre, S. S.; Mabry, J. M.; Cohen, R. E.; McKinley, G. H. Solution
42
43
Spraying of Poly(methyl Methacrylate) Blends to Fabricate Microtextured,
44 Superoleophobic Surfaces. Polymer 2011, 52 (14), 3209–3218.
45 (20) Shenoy, S. L.; Bates, W. D.; Frisch, H. L.; Wnek, G. E. Role of Chain Entanglements on
46 Fiber Formation during Electrospinning of Polymer Solutions: Good Solvent, Non-
47 Specific Polymer–polymer Interaction Limit. Polymer 2005, 46 (10), 3372–3384.
48 (21) Palangetic, L.; Reddy, N. K.; Srinivasan, S.; Cohen, R. E.; McKinley, G. H.; Clasen, C.
49
50 Dispersity and Spinnability: Why Highly Polydisperse Polymer Solutions Are Desirable
51 for Electrospinning. Polymer 2014, 55 (19), 4920–4931.
52 (22) McKinley, G. H. Dimensionless Groups for Understanding Free Surface Flows of
53 Complex Fluids. Soc Rheol Bull 2005, 2005, 6–9.
54 (23) William W. Graessley. The Entanglement Concept in Polymer Rheology; Advances in
55
56
Polymer Science; Springer Berlin Heidelberg, 1974; Vol. 16.
57
58
59
60
ACS Paragon Plus Environment
37
ACS Applied Materials & Interfaces Page 38 of 55

1
2
3
(24) Sinha-Ray, S.; Sinha-Ray, S.; Yarin, A. L.; Pourdeyhimi, B. Theoretical and Experimental
4
5 Investigation of Physical Mechanisms Responsible for Polymer Nanofiber Formation in
6 Solution Blowing. Polymer 2015, 56, 452–463.
7 (25) Lou, H.; Li, W.; Li, C.; Wang, X. Systematic Investigation on Parameters of Solution
8 Blown Micro/nanofibers Using Response Surface Methodology Based on Box-Behnken
9 Design. J. Appl. Polym. Sci. 2013, 130 (2), 1383–1391.
10
11
(26) Oliveira, J. E.; Moraes, E. A.; Costa, R. G. F.; Afonso, A. S.; Mattoso, L. H. C.; Orts, W.
12 J.; Medeiros, E. S. Nano and Submicrometric Fibers of poly(D,L-Lactide) Obtained by
13 Solution Blow Spinning: Process and Solution Variables. J. Appl. Polym. Sci. 2011, 122
14 (5), 3396–3405.
15 (27) Kolbasov, A.; Sinha-Ray, S.; Joijode, A.; Hassan, M. A.; Brown, D.; Maze, B.;
16
Pourdeyhimi, B.; Yarin, A. L. Industrial-Scale Solution Blowing of Soy Protein
17
18 Nanofibers. Ind. Eng. Chem. Res. 2016, 55 (1), 323–333.
19 (28) Zhang, L.; Kopperstad, P.; West, M.; Hedin, N.; Fong, H. Generation of Polymer Ultrafine
20 Fibers through Solution (Air-) Blowing. J. Appl. Polym. Sci. 2009, 114 (6), 3479–3486.
21 (29) Santos, A. M. C.; Medeiros, E. L. G.; Blaker, J. J.; Medeiros, E. S. Aqueous Solution
22 Blow Spinning of Poly(vinyl Alcohol) Micro- and Nanofibers. Mater. Lett. 2016, 176,
23
24
122–126.
25 (30) Oliveira, J.; Brichi, G. S.; Marconcini, J. M.; Mattoso, L. H. C.; Glenn, G. M.; Medeiros,
26 E. S. Effect of Solvent on the Physical and Morphological Properties of Poly (Lactic Acid)
27 Nanofibers Obtained by Solution Blow Spinning. J. Eng. Fibers Fabr. 2014, 9 (4), 117–
28 125.
29
(31) Nicodemo, L.; Nicolais, L. Viscosity of Bead Suspensions in Polymeric Solutions. J. Appl.
30
31 Polym. Sci. 1974, 18 (9), 2809–2818.
32 (32) Drew, C.; Wang, X.; Samuelson, L. A.; Kumar, J. The Effect of Viscosity and Filler on
33 Electrospun Fiber Morphology. J. Macromol. Sci. Part A 2003, 40 (12), 1415–1422.
34 (33) Baji, A.; Mai, Y.-W.; Wong, S.-C.; Abtahi, M.; Chen, P. Electrospinning of Polymer
35 Nanofibers: Effects on Oriented Morphology, Structures and Tensile Properties. Compos.
36
37 Sci. Technol. 2010, 70 (5), 703–718.
38 (34) Baker, B. M.; Mauck, R. L. The Effect of Nanofiber Alignment on the Maturation of
39 Engineered Meniscus Constructs. Biomaterials 2007, 28 (11), 1967–1977.
40 (35) Deitzel, J. M.; Kleinmeyer, J.; Harris, D. E. A.; Tan, N. B. The Effect of Processing
41 Variables on the Morphology of Electrospun Nanofibers and Textiles. Polymer 2001, 42
42
43
(1), 261–272.
44 (36) Gibson, P.; Schreuder-Gibson, H.; Rivin, D. Transport Properties of Porous Membranes
45 Based on Electrospun Nanofibers. Colloids Surf. Physicochem. Eng. Asp. 2001, 187, 469–
46 481.
47 (37) Behrens, A. M.; Kim, J.; Hotaling, N.; Seppala, J. E.; Kofinas, P.; Tutak, W. Rapid
48 Fabrication of poly(DL-Lactide) Nanofiber Scaffolds with Tunable Degradation for Tissue
49
50 Engineering Applications by Air-Brushing. Biomed. Mater. 2016, 11 (3), 35001. DOI:
51 10.1088/1748-6041/11/3/035001.
52 (38) Bolbasov, E. N.; Stankevich, K. S.; Sudarev, E. A.; Bouznik, V. M.; Kudryavtseva, V. L.;
53 Antonova, L. V.; Matveeva, V. G.; Anissimov, Y. G.; Tverdokhlebov, S. I. The
54 Investigation of the Production Method Influence on the Structure and Properties of the
55
56
Ferroelectric Nonwoven Materials Based on Vinylidene Fluoride – Tetrafluoroethylene
57 Copolymer. Mater. Chem. Phys. 2016, 182, 338–346.
58
59
60
ACS Paragon Plus Environment
38
Page 39 of 55 ACS Applied Materials & Interfaces

1
2
3
(39) Oliveira, J. E.; Mattoso, L. H.; Orts, W. J.; Medeiros, E. S. Structural and Morphological
4
5 Characterization of Micro and Nanofibers Produced by Electrospinning and Solution Blow
6 Spinning: A Comparative Study. Adv. Mater. Sci. Eng. 2013, 409572.
7 (40) Fraley, S. I.; Wu, P.; He, L.; Feng, Y.; Krisnamurthy, R.; Longmore, G. D.; Wirtz, D.
8 Three-Dimensional Matrix Fiber Alignment Modulates Cell Migration and MT1-MMP
9 Utility by Spatially and Temporally Directing Protrusions. Sci. Rep. 2015, 5, 14580. DOI:
10
11
10.1038/srep14580.
12 (41) Medeiros, E. L. G.; Braz, A. L.; Porto, I. J.; Menner, A.; Bismarck, A.; Boccaccini, A. R.;
13 Lepry, W. C.; Nazhat, S. N.; Medeiros, E. S.; Blaker, J. J. Porous Bioactive Nanofibers via
14 Cryogenic Solution Blow Spinning and Their Formation into 3D Macroporous Scaffolds.
15 ACS Biomater. Sci. Eng. 2016, 2 (9), 1442–1449.
16
(42) Bhattarai, N.; Edmondson, D.; Veiseh, O.; Matsen, F. A.; Zhang, M. Electrospun
17
18 Chitosan-Based Nanofibers and Their Cellular Compatibility. Biomaterials 2005, 26 (31),
19 6176–6184.
20 (43) Teo, W. E.; Ramakrishna, S. A Review on Electrospinning Design and Nanofibre
21 Assemblies. Nanotechnology 2006, 17 (14), R89–R106.
22 (44) Wu, H.; Zhang, R.; Liu, X.; Lin, D.; Pan, W. Electrospinning of Fe, Co, and Ni
23
24
Nanofibers: Synthesis, Assembly, and Magnetic Properties. Chem. Mater. 2007, 19 (14),
25 3506–3511.
26 (45) Li, D.; Wang, Y.; Xia, Y. Electrospinning of Polymeric and Ceramic Nanofibers as
27 Uniaxially Aligned Arrays. Nano Lett. 2003, 3 (8), 1167–1171.
28 (46) Lou, H.; Han, W.; Wang, X. Numerical Study on the Solution Blowing Annular Jet and Its
29
Correlation with Fiber Morphology. Ind. Eng. Chem. Res. 2014, 53 (7), 2830–2838.
30
31 (47) Collet, J.-P.; Shuman, H.; Ledger, R. E.; Lee, S.; Weisel, J. W. The Elasticity of an
32 Individual Fibrin Fiber in a Clot. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (26), 9133–
33 9137.
34 (48) Agache, P.G.; Monneur, C.; Leveque, J.L.; De Rigal, J. Mechanical Properties and
35 Young’s Modulus of Human Skin in Vivo. Arch. Dermatol. Res. 1980, 269 (3), 221–232.
36
37 (49) Oliveira, J. E.; Moraes, E. A.; Marconcini, J. M.; C. Mattoso, L. H.; Glenn, G. M.;
38 Medeiros, E. S. Properties of Poly(lactic Acid) and Poly(ethylene Oxide) Solvent Polymer
39 Mixtures and Nanofibers Made by Solution Blow Spinning. J. Appl. Polym. Sci. 2013, 129
40 (6), 3672–3681.
41 (50) Zhuang, X.; Yang, X.; Shi, L.; Cheng, B.; Guan, K.; Kang, W. Solution Blowing of
42
43
Submicron-Scale Cellulose Fibers. Carbohydr. Polym. 2012, 90 (2), 982–987.
44 (51) Sinha-Ray, S.; Yarin, A. L.; Pourdeyhimi, B. The Production of 100/400 Nm Inner/outer
45 Diameter Carbon Tubes by Solution Blowing and Carbonization of Core–shell
46 Nanofibers. Carbon 2010, 48 (12), 3575–3578.
47 (52) Sinha-Ray, S.; Zhang, Y.; Yarin, A. L.; Davis, S. C.; Pourdeyhimi, B. Solution Blowing of
48 Soy Protein Fibers. Biomacromolecules 2011, 12 (6), 2357–2363.
49
50 (53) Lee, M. W.; Yoon, S. S.; Yarin, A. L. Solution-Blown Core–Shell Self-Healing Nano- and
51 Microfibers. ACS Appl. Mater. Interfaces 2016, 8 (7), 4955–4962.
52 (54) Lee, M. W.; Sett, S.; Yoon, S. S.; Yarin, A. L. Fatigue of Self-Healing Nanofiber-Based
53 Composites: Static Test and Subcritical Crack Propagation. ACS Appl. Mater. Interfaces
54 2016, 8 (28), 18462–18470.
55
56
57
58
59
60
ACS Paragon Plus Environment
39
ACS Applied Materials & Interfaces Page 40 of 55

1
2
3
(55) Sinha-Ray, S.; Pelot, D. D.; Z. P. Zhou; Rahman, A.; Wu, X.-F.; Yarin, A. L.
4
5 Encapsulation of Self-Healing Materials by Coelectrospinning, Emulsion Electrospinning,
6 Solution Blowing and Intercalation. J. Mater. Chem. 2012, 22 (18), 9138-9146.
7 (56) Liu, R.; Xu, X.; Zhuang, X.; Cheng, B. Solution Blowing of chitosan/PVA Hydrogel
8 Nanofiber Mats. Carbohydr. Polym. 2014, 101, 1116–1121.
9 (57) Behrens, A. M.; Lee, N. G.; Casey, B. J.; Srinivasan, P.; Sikorski, M. J.; Daristotle, J. L.;
10
11
Sandler, A. D.; Kofinas, P. Biodegradable-Polymer-Blend-Based Surgical Sealant with
12 Body-Temperature-Mediated Adhesion. Adv. Mater. 2015, 27 (48), 8056–8061.
13 (58) Xu, X.; Zhou, G.; Li, X.; Zhuang, X.; Wang, W.; Cai, Z.; Li, M.; Li, H. Solution Blowing
14 of chitosan/PLA/PEG Hydrogel Nanofibers for Wound Dressing. Fibers Polym. 2016, 17
15 (2), 205–211.
16
(59) Hoffman, K.; Skrtic, D.; Sun, J.; Tutak, W. Airbrushed Composite Polymer Zr-ACP
17
18 Nanofiber Scaffolds with Improved Cell Penetration for Bone Tissue Regeneration. Tissue
19 Eng. Part C Methods 2014, 21 (3), 284–291.
20 (60) Wang, J.; Jákli, A.; West, J. L. Airbrush Formation of Liquid Crystal/Polymer Fibers.
21 ChemPhysChem 2015, 16 (9), 1839–1841.
22 (61) Tutak, W.; Jyotsnendu, G.; Bajcsy, P.; Simon, C. G. Nanofiber Scaffolds Influence
23
24
Organelle Structure and Function in Bone Marrow Stromal Cells. J. Biomed. Mater. Res.
25 B Appl. Biomater. 2016. [Online early access] DOI: 10.1002/jbm.b.33624.
26 (62) Litviakov, N. V.; Tverdokhlebov, S. I.; Perelmuter, V. M.; Kulbakin, D. E.; Bolbasov, E.
27 N.; Tsyganov, M. M.; Zheravin, A. A.; Svetlichnyi, V. A.; Cherdyntseva, N. V. Composite
28 Implants Coated with Biodegradable Polymers Prevent Stimulating Tumor Progression;
29
2016; p 20043.
30
31 (63) Billiet, T.; Vandenhaute, M.; Schelfhout, J.; Van Vlierberghe, S.; Dubruel, P. A Review of
32 Trends and Limitations in Hydrogel-Rapid Prototyping for Tissue Engineering.
33 Biomaterials 2012, 33 (26), 6020–6041.
34 (64) Dalton, P. D.; Klinkhammer, K.; Salber, J.; Klee, D.; Möller, M. Direct in Vitro
35 Electrospinning with Polymer Melts. Biomacromolecules 2006, 7 (3), 686–690.
36
37 (65) Lu, B.; He, Y.; Duan, H.; Zhang, Y.; Li, X.; Zhu, C.; Xie, E. A New Ultrahigh-Speed
38 Method for the Preparation of Nanofibers Containing Living Cells: A Bridge towards
39 Industrial Bioengineering Applications. Nanoscale 2012, 4 (3), 1003-1009.
40 (66) Yuan, S.; Li, Z.; Song, L.; Shi, H.; Luan, S.; Yin, J. Liquid-Infused Poly(styrene- B -
41 Isobutylene- B -Styrene) Microfiber Coating Prevents Bacterial Attachment and
42
43
Thrombosis. ACS Appl. Mater. Interfaces 2016, 8 (33), 21214–21220.
44 (67) Bilbao-Sainz, C.; Chiou, B.-S.; Valenzuela-Medina, D.; Du, W.-X.; Gregorski, K. S.;
45 Williams, T. G.; Wood, D. F.; Glenn, G. M.; Orts, W. J. Solution Blow Spun Poly(lactic
46 Acid)/hydroxypropyl Methylcellulose Nanofibers with Antimicrobial Properties. Eur.
47 Polym. J. 2014, 54, 1–10.
48 (68) Bonan, R. F.; Bonan, P. R. F.; Batista, A. U. D.; Sampaio, F. C.; Albuquerque, A. J. R.;
49
50 Moraes, M. C. B.; Mattoso, L. H. C.; Glenn, G. M.; Medeiros, E. S.; Oliveira, J. E. In
51 Vitro Antimicrobial Activity of Solution Blow Spun Poly(lactic
52 Acid)/polyvinylpyrrolidone Nanofibers Loaded with Copaiba (Copaifera Sp.) Oil. Mater.
53 Sci. Eng. C 2015, 48, 372–377.
54 (69) Oliveira, J. E.; Medeiros, E. S.; Cardozo, L.; Voll, F.; Madureira, E. H.; Mattoso, L. H. C.;
55
56
Assis, O. B. G. Development of Poly(lactic Acid) Nanostructured Membranes for the
57
58
59
60
ACS Paragon Plus Environment
40
Page 41 of 55 ACS Applied Materials & Interfaces

1
2
3
Controlled Delivery of Progesterone to Livestock Animals. Mater. Sci. Eng. C 2013, 33
4
5 (2), 844–849.
6 (70) Zhuang, X.; Shi, L.; Zhang, B.; Cheng, B.; Kang, W. Coaxial Solution Blown Core-Shell
7 Structure Nanofibers for Drug Delivery. Macromol. Res. 2013, 21 (4), 346–348.
8 (71) Khansari, S.; Duzyer, S.; Sinha-Ray, S.; Hockenberger, A.; Yarin, A. L.; Pourdeyhimi, B.
9 Two-Stage Desorption-Controlled Release of Fluorescent Dye and Vitamin from Solution-
10
11
Blown and Electrospun Nanofiber Mats Containing Porogens. Mol. Pharm. 2013, 10 (12),
12 4509–4526.
13 (72) Oliveira, J. E.; Mattoso, L. H. C.; Medeiros, E. S.; Zucolotto, V. Poly(lactic acid)/Carbon
14 Nanotube Fibers as Novel Platforms for Glucose Biosensors. Biosensors 2012, 2 (4), 70–
15 82.
16
(73) Oliveira, J. E.; Grassi, V.; Scagion, V. P.; Mattoso, L. H. C.; Glenn, G. M.; Medeiros, E.
17
18 S. Sensor Array for Water Analysis Based on Interdigitated Electrodes Modified With
19 Fiber Films of Poly(Lactic Acid)/Multiwalled Carbon Nanotubes. IEEE Sens. J. 2013, 13
20 (2), 759–766.
21 (74) Manufactured Fibre Technology; Gupta, V. B., Kothari, V. K., Eds.; Springer
22 Netherlands: Dordrecht, 1997.
23
24
(75) Tuteja, A.; Choi, W.; Ma, M.; Rutledge, G. C.; McKinley, G. H.; Cohen, R. E.; Mazzella,
25 S. A.; Mabry, J. M. Designing Superoleophobic Surfaces (Postprint); DTIC Document,
26 2007.
27 (76) Shi, L.; Zhuang, X.; Tao, X.; Cheng, B.; Kang, W. Solution Blowing Nylon 6 Nanofiber
28 Mats for Air Filtration. Fibers Polym. 2013, 14 (9), 1485–1490.
29
(77) Lee, J.-G.; Kim, D.-Y.; Mali, M. G.; Al-Deyab, S. S.; Swihart, M. T.; Yoon, S. S.
30
31 Supersonically Blown Nylon-6 Nanofibers Entangled with Graphene Flakes for Water
32 Purification. Nanoscale 2015, 7 (45), 19027–19035.
33 (78) Hsiao, H.-Y.; Huang, C.-M.; Hsu, M.-Y.; Chen, H. Preparation of High-Surface-Area
34 PAN-Based Activated Carbon by Solution-Blowing Process for CO2 Adsorption. Sep.
35 Purif. Technol. 2011, 82, 19–27.
36
37 (79) Tao, X.; Zhou, G.; Zhuang, X.; Cheng, B.; Li, X.; Li, H. Solution Blowing of Activated
38 Carbon Nanofibers for Phenol Adsorption. RSC Adv 2015, 5 (8), 5801–5808.
39 (80) Sinha-Ray, S.; Sinha-Ray, S.; Yarin, A. L.; Pourdeyhimi, B. Application of Solution-
40 Blown 20–50 Nm Nanofibers in Filtration of Nanoparticles: The Efficient van Der Waals
41 Collectors. J. Membr. Sci. 2015, 485, 132–150.
42
43
(81) Wang, H.; Zhuang, X.; Li, X.; Wang, W.; Wang, Y.; Cheng, B. Solution Blown
44 Sulfonated Poly(ether Sulfone)/poly(ether Sulfone) Nanofiber-Nafion Composite
45 Membranes for Proton Exchange Membrane Fuel Cells. J. Appl. Polym. Sci. 2015, 132
46 (38), 42572.
47 (82) Wang, H.; Zhuang, X.; Tong, J.; Li, X.; Wang, W.; Cheng, B.; Cai, Z. Solution-Blown
48 SPEEK/POSS Nanofiber–nafion Hybrid Composite Membranes for Direct Methanol Fuel
49
50 Cells. J. Appl. Polym. Sci. 2015, 132 (47), 42843. DOI: 10.1002/app.42843.
51 (83) Xu, X.; Li, L.; Wang, H.; Li, X.; Zhuang, X. Solution Blown Sulfonated Poly(ether Ether
52 Ketone) nanofiber–Nafion Composite Membranes for Proton Exchange Membrane Fuel
53 Cells. RSC Adv 2015, 5 (7), 4934–4940.
54 (84) Shi, S.; Zhuang, X.; Cheng, B.; Wang, X. Solution Blowing of ZnO Nanoflake-
55
56
Encapsulated Carbon Nanofibers as Electrodes for Supercapacitors. J. Mater. Chem. A
57 2013, 1 (44), 13779.
58
59
60
ACS Paragon Plus Environment
41
ACS Applied Materials & Interfaces Page 42 of 55

1
2
3
(85) Jia, K.; Zhuang, X.; Cheng, B.; Shi, S.; Shi, Z.; Zhang, B. Solution Blown Aligned Carbon
4
5 Nanofiber Yarn as Supercapacitor Electrode. J. Mater. Sci. Mater. Electron. 2013, 24 (12),
6 4769–4773.
7 (86) Li, L.; Kang, W.; Zhao, Y.; Li, Y.; Shi, J.; Cheng, B. Preparation of Flexible Ultra-Fine
8 Al2O3 Fiber Mats via the Solution Blowing Method. Ceram. Int. 2015, 41 (1), 409–415.
9 (87) Cheng, B.; Tao, X.; Shi, L.; Yan, G.; Zhuang, X. Fabrication of ZrO2 Ceramic Fiber Mats
10
11
by Solution Blowing Process. Ceram. Int. 2014, 40 (9), 15013–15018.
12 (88) Farias, R. M. da C.; Menezes, R. R.; Oliveira, J. E.; de Medeiros, E. S. Production of
13 Submicrometric Fibers of Mullite by Solution Blow Spinning (SBS). Mater. Lett. 2015,
14 149, 47–49.
15 (89) Rotta, M.; Zadorosny, L.; Carvalho, C. L.; Malmonge, J. A.; Malmonge, L. F.; Zadorosny,
16
R. YBCO Ceramic Nanofibers Obtained by the New Technique of Solution Blow
17
18 Spinning. Ceram. Int. 2016, 42 (14), 16230–16234.
19 (90) Costa, D. L.; Leite, R. S.; Neves, G. A.; Santana, L. N. de L.; Medeiros, E. S.; Menezes,
20 R. R. Synthesis of TiO2 and ZnO Nano and Submicrometric Fibers by Solution Blow
21 Spinning. Mater. Lett. 2016, 183, 109–113.
22 (91) Gonzalez-Abrego, M.; Hernandez-Granados, A.; Guerrero-Bermea, C.; Martinez de la
23
24
Cruz, A.; Garcia-Gutierrez, D.; Sepulveda-Guzman, S.; Cruz-Silva, R. Mesoporous
25 Titania Nanofibers by Solution Blow Spinning. J. Sol-Gel Sci. Technol. 2016.
26 (92) Giannatsis, J.; Dedoussis, V. Additive Fabrication Technologies Applied to Medicine and
27 Health Care: A Review. Int. J. Adv. Manuf. Technol. 2009, 40 (1–2), 116–127.
28 (93) Zeng, W.; Shu, L.; Li, Q.; Chen, S.; Wang, F.; Tao, X.-M. Fiber-Based Wearable
29
Electronics: A Review of Materials, Fabrication, Devices, and Applications. Adv. Mater.
30
31 2014, 26 (31), 5310–5336.
32 (94) Tutak, W.; Kaufman, G.; Gelven, G.; Markle, C.; Maczka, C. Uniform, Fast, High
33 Concentration Delivery of Bone Marrow Stromal Cells and Gingival Fibroblasts by Gas-
34 Brushing. Biomed. Phys. Eng. Express 2016, 2 (3), 35007.
35 (95) Veazey, W. S.; Anusavice, K. J.; Moore, K. Mammalian Cell Delivery via Aerosol
36
37 Deposition. J. Biomed. Mater. Res. B Appl. Biomater. 2005, 72B (2), 334–338.
38 (96) Pehlivaner Kara, M. O.; Ekenseair, A. K. In Situ Spray Deposition of Cell-Loaded,
39 Thermally and Chemically Gelling Hydrogel Coatings for Tissue Regeneration. J.
40 Biomed. Mater. Res. A 2016, 104 (10), 2383–2393.
41 (97) Demirel, M. C.; Cetinkaya, M.; Pena-Francesch, A.; Jung, H. Recent Advances in
42
43
Nanoscale Bioinspired Materials: Recent Advances in Nanoscale Bioinspired Materials.
44 Macromol. Biosci. 2015, 15 (3), 300–311.
45 (98) Khansari, S.; Sinha-Ray, S.; Yarin, A. L.; Pourdeyhimi, B. Biopolymer-Based Nanofiber
46 Mats and Their Mechanical Characterization. Ind. Eng. Chem. Res. 2013, 52 (43), 15104–
47 15113.
48 (99) Sett, S.; Stephansen, K.; Yarin, A. L. Solution-Blown Nanofiber Mats from Fish
49
50 Sarcoplasmic Protein. Polymer 2016, 93, 78–87.
51 (100) da Silva Parize, D. D.; de Oliveira, J. E.; Foschini, M. M.; Marconcini, J. M.; Mattoso, L.
52 H. C. Poly(lactic Acid) Fibers Obtained by Solution Blow Spinning: Effect of a Greener
53 Solvent on the Fiber Diameter. J. Appl. Polym. Sci. 2016, 133 (18), 43379.
54 (101) Zhuang, X.; Shi, L.; Jia, K.; Cheng, B.; Kang, W. Solution Blown Nanofibrous Membrane
55
56
for Microfiltration. J. Membr. Sci. 2013, 429, 66–70.
57
58
59
60
ACS Paragon Plus Environment
42
Page 43 of 55 ACS Applied Materials & Interfaces

1
2
3
(102) Mahalingam, S.; Edirisinghe, M. Forming of Polymer Nanofibers by a Pressurised
4
5 Gyration Process. Macromol. Rapid Commun. 2013, 34 (14), 1134–1139.
6 (103) Yetisen, A. K.; Qu, H.; Manbachi, A.; Butt, H.; Dokmeci, M. R.; Hinestroza, J. P.;
7 Skorobogatiy, M.; Khademhosseini, A.; Yun, S. H. Nanotechnology in Textiles. ACS
8 Nano 2016, 10 (3), 3042–3068.
9 (104) Tutak, W.; Gelven, G.; Markle, C.; Palmer, X.-L. Rapid Polymer Fiber Airbrushing:
10
11
Impact of a Device Design on the Fiber Fabrication and Matrix Quality. J. Appl. Polym.
12 Sci. 2015, 132 (47), 42813. DOI: 10.1002/app.42813.
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
43
ACS Applied Materials & Interfaces Page 44 of 55

1
2
3
TABLE OF CONTENTS IMAGE
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
44
Page 45 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Figure 1. A) Schematic for a solution blow spinning device showing concentric nozzles. Adapted with
35 permission.[1] Copyright 2015, American Chemical Society. B) A general solution blow spinning process
36 diagram with rolling collector. Adapted with permission.[3] Copyright 2009, Wiley-VCH. C) Image of direct
37 deposition of poly(styrene-block-isoprene-block-styrene) block copolymer fibers using a homemade solution
38 blow spinning device. The homemade device was made from a transfer pipet and a flat tipped 18G needle
39 (inset). Adapted with permission.[1] Copyright 2015, American Chemical Society. D) A commercial airbrush
40 used for solution blow spinning. Adapted with permission.[2] Copyright 2013, Elsevier.
41 Figure 1A
435x358mm (300 x 300 DPI)
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 46 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 2. The number of publications on solution blow spinning and related topics has increased since its
29 widespread exposure began in 2009. Data for 2016 is current as of September.
30 Figure 2
279x172mm (300 x 300 DPI)
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 47 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Figure 3. A) Plot indicating morphology of poly(methyl methacrylate) (PMMA) sprayed using a solution blow
32
spinning (SBS) apparatus at various concentrations and molecular weights. The estimated overlap
33 concentration (c*) is indicated by the dashed line. Scanning electron microscopy (SEM) image of PMMA
34 fibers formed at high molecular weight but below overlap concentration. Scale bar represents 50 µm. B)
35 SEM images of 50/50 wt.% PMMA/1H,1H,2H,2H-heptadecafluorodecyl polyhedral oligomeric silsesquioxane
36 (PMMA: Mw = 593 kDa, PDI = 2.69) blends sprayed using an SBS apparatus at increasing concentrations of
37 PMMA in solution. Scale bars represent 50 µm, 100 µm, and 50 µm, respectively. Adapted with
38 permission.[19] Copyright 2011, Elsevier.
39 Figure 3A
177x130mm (300 x 300 DPI)
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 48 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 Figure 4. A–C) Scanning electron microscopy (SEM) images of fiber morphologies produced by solution blow
16 spinning 10% w/v poly(lactic-co-glycolic acid) (PLGA) in acetone at CO2 flow rates of 10, 13, and 15 SCFH.
17 Scale bars correspond to 20 µm. Adapted with permission.[5] Copyright 2014, American Chemical Society.
18 Figure 4A
19 507x111mm (300 x 300 DPI)
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 49 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 5. Morphology, Young’s modulus, and cyclic testing of solution blow spinning (SBS) fibers. A) Fiber
31 bundles produced by SBS of a 4 wt.% polymer solution of poly(caprolactone) (PCL) (80 kDa) in chloroform.
B) Fibers produced by electrospinning a 4 wt.% polymer solution of PCL (80 kDa) in 3:1
32 chloroform:methanol by mass. A-B adapted with permission.[2] Copyright 2013, Elsevier. C) Comparison of
33 stress-strain curves for PCL (80000 kDa) fiber mats fabricated using electrospinning and SBS. D) Young’s
34 modulus for fiber mats produced using SBS and electrospinning PCL fibers. C–D adapted with permission.[2]
35 Copyright 2013, Elsevier. E) Stress/strain cycling curves at varying maximum strain values for poly(styrene-
36 block-isoprene-block-styrene) fiber mats fabricated by SBS. Adapted with permission.[1] Copyright 2015,
37 American Chemical Society.
38 Figure 5A
555x374mm (300 x 300 DPI)
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 50 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
Figure 6. Images showing the effects of turbulence on polymer jets used for solution blow spinning (SBS).
47
A) High-speed photograph of a polymer solution cone forming at the tip of a SBS device. Adapted with
48 permission.[3] Copyright 2009, Wiley-VCH. B–C) High-speed photograph comparing the length of the
49 straight region (in red box) of the polymer jet produced by SBS at gas pressures of 1.121 atm (B) and
50 1.363 atm (C). Adapted with permission.[40] Copyright 2014, American Chemical Society.
51 Figure 6A
52 215x260mm (300 x 300 DPI)
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 51 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
Figure 7. Various multicomponent fibers produced by solution blow spinning (SBS). A) Transmission electron
30 microscopy images of core-shell fibers produced by SBS from a 1:1 blend of poly(lactic acid) and
31 poly(ethylene glycol) (PEG) at 6% w/v in solution. Adapted with permission.[49] Copyright 2013, Wiley-
32 VCH. B–C) Scanning electron microscopy (SEM) images of fibers produced by SBS from a 1:1 blend of
33 chitosan and poly(vinyl alcohol) at 8% w/v in solution with 7% ethylene glycol diglycidyl ether crosslinker
34 (relative to mass of polymer). SEM images were taken after drying in a vacuum oven for 3 hours (B), and
35 after drying then swelling in saline solution for 8 hours (C). Adapted with permission.[56] Copyright 2014,
36 Elsevier. D) Fibers produced by SBS from a 10% w/v poly(lactic-co-glycolic acid) (PLGA) in acetone solution.
E) Fibers produced by SBS from a 10% w/v PLGA, 5% w/v PEG blend in solution. F) Fiber diameter
37
produced by SBS from a 10% w/v PLGA solution increases with increasing PEG blend content. Adapted with
38 permission.[57] Copyright 2015, Wiley-VCH.
39 Figure 7A
40 376x251mm (300 x 300 DPI)
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 52 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 Figure 8. Biomaterials applications of solution blow spinning (SBS). A–C) Fluorescent microscopy images
28 showing MSCs cultured on a flat cell culture control surface (A), polyvinylidene fluoride-trifluroethylene
29 copolymer (VDF-TeFE) fibers produced by electrospinning (B), and VDF-TeFE fibers produced by SBS (C).
30 Scale bars represent 100 µm. Adapted with permission.[36] Copyright 2016, Elsevier. D–E) Field emission
31 scanning electron microscopy (FESEM) images showing blood cell and platelet attachment to poly(styrene-b-
32 isobutylene-b-styrene) (SIBS) fibers (D) and SIBS fibers infused with silicon oil (E) after incubation with
33 whole blood. Scale bars represent 10 µm. Adapted with permission.[60] Copyright 2016, American Chemical
Society. F) Intestinal anastomosis being coated in fibers produced by SBS from a poly(lactic-co-glycolic acid)
34 (PLGA) and poly(ethylene glycol) (PEG) blend. G) Intestinal anastomosis 2 minutes after being coated
35 showing the thermal transition of PLGA-PEG fibers. F–G adapted with permission.[51] Copyright 2015,
36 Wiley-VCH.
37 Figure 8A
38 344x207mm (300 x 300 DPI)
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 53 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17 Figure 9. Composite and post-processing techniques enabled by solution blow spinning (SBS). A) Glove
18 coated in poly(styrene-block-isoprene-block-styrene) elastomeric fibers patterned with conductive bands of
19 silver nanoparticles showing resistance measured across each line corresponding to hand gestures. Adapted
20 with permission.[1] Copyright 2015, American Chemical Society. B–C) Side (B) and cross section (C) of
21 carbon tubes processed from core-shell poly(methyl methacrylate)-poly(acrylonitrile) fibers produced by
SBS. Adapted with permission.[45] Copyright 2010, Elsevier.
22
Figure 9A
23 487x137mm (300 x 300 DPI)
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 54 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Figure 10. Future directions for solution blow spinning (SBS) research. A–B) Images of nanofibers produced
19
by SBS from a 15% polyvinylpyrrolidone solution in fetal bovine serum sprayed without (A) and with (B)
20 mouse renal epithelial cells in the solution. Adapted with permission.[59] Copyright 2012, RSC Publishing.
21 C) Multinozzle SBS process using a die with 41 nozzles per row and 8 rows. Adapted with permission.[27]
22 Copyright 2016, American Chemical Society.
23 Figure 10A
24 456x148mm (300 x 300 DPI)
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 55 of 55 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Table of Contents Image
33
34 273x204mm (300 x 300 DPI)
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

You might also like