Sie sind auf Seite 1von 8

Hollow core photonic crystal fibers for beam

delivery
G. Humbert, J. C. Knight, G. Bouwmans, P. St. J. Russell
Department of Physics, University of Bath, Claverton Down, Bath, BA2 7AY, United Kingdom
j.c.knight@bath.ac.uk

D. P. Williams, P. J. Roberts, B. J. Mangan


BlazePhotonics Ltd., University of Bath Campus, Claverton Down, Bath, BA2 7AY, United Kingdom

Abstract: Hollow-core photonic crystal fibers have unusual properties


which make them ideally suited to delivery of laser beams. We describe the
properties of fibers with different core designs, and the observed effects of
anti-crossings with interface modes. We conclude that 7-unit-cell cores are
currently most suitable for transmission of femtosecond and sub-picosecond
pulses, whereas larger cores (e.g. 19-cell cores) are better for delivering
nanosecond pulsed and continuous-wave beams.
2004 Optical Society of America
OCIS codes: (060.2310) Fiber Optics, (160.2290) Fiber Materials.

References and links


1. J. C. Knight, “Photonic Crystal fibers,” Nature 424, 847-851 (2003).
2. B. J. Mangan, L. Farr, A. Langford, P. J. Roberts, D. P. Williams, F. Couny, M. Lawman, M. Mason, S.
Coupland, R. Flea, H. Sabert, T. A. Birks, J. C. Knight, P. St.J. Russell, “Low loss (1.7 dB/km) hollow core
photonic bandgap fiber,” postdeadline paper PDP24, OFC’04 (Los Angeles, 2004).
3. P. St.J. Russell, “Photonic crystal fibres,” Science 299, 358-362 (2003).
4. D. G. Ouzounov, F. R. Ahmad, D. Muller, N. Venkataraman, M. T. Gallagher, M. G. Thomas, J. Silcox, K.W.
Koch, A. L. Gaeta, “Generation of megawatt optical solitons in hollow-core photonic band-gap fibers,” Science
301 1702-1704 (2003).
5. J. D. Shephard, J. D. C. Jones, D. P. Hand, G. Bouwmans, J. C. Knight, P. St.J. Russell, B. J. Mangan, "High
energy nanosecond laser pulses delivered single-mode through hollow-core PBG fibers," Opt. Express 12, 717-
723 (2004), http://www.opticsexpress.org/abstract.cfm?URI=OPEX-12-4-717
6. R. F. Cregan, B. J. Mangan, J. C. Knight, T. A. Birks, P. St.J. Russell, D. Allen, P. J. Roberts, “Single-mode
photonic bandgap guidance of light in air,” Science 285, 1537-1539 (1999).
7. C. M. Smith, N. Venkataraman, M. T. Gallagher, D. Muller, J. A. West, N. F. Borrelli, D. C. Allan, K. Koch,
“Low-loss hollow-core silica/air photonic bandgap fibre,” Nature 424, 657-659 (2003).
8. D. C. Allan, N. F. Borrelli, M. T. Gallagher, D. Müller, C. M. Smith, N. Venkataraman, J. A. West, P. Zhang
and K. W. Koch, “Surface modes and loss in air-core photonic band-gap fibers”, in Photonic Crystal Materials
and Devices, A. Adibi, A. Scherer, S. Y. Lin, Eds. Proc. SPIE 5000 (2003)
9. F. Luan, J. C. Knight, P. S. J. Russell, S. Campbell, D. Xiao, D. T. Reid, B. J. Mangan, D. P. Williams, and
P. J. Roberts, "Femtosecond soliton pulse delivery at 800nm wavelength in hollow-core photonic bandgap
fibers," Opt. Express 12, 835-840 (2004), http://www.opticsexpress.org/abstract.cfm?URI=OPEX-12-5-835
10. K. Saitoh, N. A. Mortensen, and M. Koshiba, "Air-core photonic band-gap fibers: the impact of surface
modes," Opt. Express 12, 394-400 (2004), http://www.opticsexpress.org/abstract.cfm?URI=OPEX-12-3-394
11. W. J. Wadsworth, N. Joly, J. C. Knight, T. A. Birks, F. Biancalana, and P. St.J. Russell, "Supercontinuum and
four-wave mixing with Q-switched pulses in endlessly single-mode photonic crystal fibres," Opt. Express 12,
299-309 (2004) http://www.opticsexpress.org/abstract.cfm?URI=OPEX-12-2-299
12. C. J. S. de Matos and J. R. Taylor, “Multi-kilowatt, all-fiber integrated chirped-pulse amplification system
yielding 40× pulse compression using air-core fiber and conventional erbium-doped fiber amplifier,” Opt.
Express 12, 405-409 (2004), http://www.opticsexpress.org/abstract.cfm?URI=OPEX-12-3-405

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1477
1. Introduction
Hollow core photonic crystal fibers have become the most advanced manifestation of 2-
dimensional photonic bandgap structures [1], enabling the guidance of light in a hollow core
with low attenuation on kilometer length scales [2] – something that is completely impossible
in conventional optical fibres. Their remarkable properties, which were studied academically
for several years after their first conception in 1991 [3], are now manifest in a number of
structures which are rapidly becoming commercially available. For these fibers to have a
lasting impact in the world of optics and beyond, it is vital that their potential for real-world
applications is recognized and realized over the next few years. One possible application area
for hollow-core photonic crystal fibers (HC-PCF’s) is in telecommunications. HC-PCF’s
could conceivably demonstrate lower, or even far lower, optical attenuation than conventional
fibers, which are limited by the optical properties of their solid cores. Another application
area, which is perhaps closer to fruition, is the field of laser delivery. With their greatly
reduced nonlinearity [4] and increased damage thresholds [5], and with dispersion
characteristics very different to conventional fibers, this application area looks like a
significant opportunity. In this paper we investigate the basic properties of HC-PCF’s which
make them suitable for delivery of continuous wave (CW), nanosecond and sub-picosecond
laser beams.
High power lasers (including CW, Q-switched and mode-locked configurations) are
widely used in fields as diverse as marking, machining and welding, laser-Doppler
velocimetry, laser surgery and THz generation. For many applications, optical fiber would be
the preferred means of delivery if it were reliable and efficient, but is currently unusable.
Delivery of powerful laser light (both pulsed and CW) using optical fibers is traditionally
hard, because the high optical powers and energies cause fiber damage and deleterious
nonlinear-optical phenomena, and also because of the group-velocity dispersion (GVD) in
fibers which disperses short pulses. Our attention here is focused especially on those
applications requiring fiber delivery with a high beam quality, such as micromachining or
laser beams for guide stars. For such applications, fiber attenuation is generally not a limiting
parameter, as the lengths involved are typically of the order of a few m. Instead, the limiting
parameters are dispersion, nonlinearity and fiber damage. In this paper we describe how these
parameters depend on the fiber design and which fibers are currently most useful for such
applications.
2. Hollow-core photonic crystal fibers
HC-PCF’s are very different to conventional optical fibers, and can be formed in several ways
using different materials. The fibers discussed in this paper are drawn from a preform created
by stacking together hundreds of hollow silica capillaries, so that the final fiber has a 2-
dimensional pattern of air holes running down its length. An electron micrograph of a HC-
PCF is shown in Fig. 1. The periodic “holey” cladding surrounds a larger central air hole,
which serves as the fiber core. Light is trapped in the core by the photonic bandgap of the
cladding, which covers a finite frequency range, typically around 20% of the central
frequency. Outside of the bandgap, the modes of the core are not confined, and the attenuation
of the fiber is high. Within the bandgap, one or more modes are localized within the vicinity
of the core, and the attenuation falls when the fundamental guided mode is most strongly
peaked in the hollow core [2,6]. The minimum attenuation is determined by scattering due to
imperfections in the fiber and by coupling to other confined modes (interface or surface
modes) which have a greater overlap with the core surround [7,8]. The properties of these
interface modes are strongly dependent on the fiber design, but almost invariably influence
the performance of the fiber as a whole. In particular, the lowest-loss wavelength is not
necessarily determined by the center of the bandgap, but by the locations of the interface
modes. Most of the fibers which have been reported have been based on 7-cell or 19-cell core
designs, formed by omitting 7 or 19 central capillaries from the stack when the preform is
being built (the fiber in Fig. 1 is a 7-cell design). In this paper we describe the observed

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1478
Fig. 1. Scanning electron micrograph of the HC-PCF used in the experiments described in this
paper

effects of interface modes and other features of these two core designs that impact on their
suitability for laser delivery.
3. Guided modes
The relevant properties of a HC-PCF – the dispersion, nonlinearity and damage threshold –
are closely related to the modal intensity patterns. Over much of the operating wavelength
range, HC-PCF’s have a quasi-Gaussian fundamental-mode field distribution, strongly peaked
in the center of the hollow core and decaying rapidly as one approaches the core walls.
However, the very different optical response of the hollow core and of the silica strands
forming the core surround mean that the small residual overlap of the guided mode with the
solid material can substantially determine the optical characteristics of the fiber. For example,
we have recently found using numerical modeling that the nonlinear phase shift in 7-cell HC-
PCF can be caused more by the nonlinear response of the solid silica than by the air in the
core, despite the overlap with the solid being just a few percent of the energy of the guided
mode [9]. In addition, the modal field patterns are radically different in the frequency range
around an anti-crossing with an interface mode [7], with a significant increase in the light-in-
glass fraction. This has a dramatic effect on the modal field profiles, the attenuation and the
GVD [10] and must be expected to alter the damage threshold as well.
The fiber studied in this work is that shown in Fig. 1. The optical attenuation of the fiber
(shown in the inset to Fig. 2) was measured with a broadband light source using the cutback
technique. The attenuation minimum was found to be 78 dB/km at a wavelength of 1102 nm,
and the main low-loss window is roughly 115 nm wide, centered at 1083 nm. In order to
identify features in the high-loss regions surrounding the transmission window, we present in
Fig. 2 the transmitted spectrum of a 5 m fiber length measured using a broadband source
(tungsten-halogen lamp). On the long-wavelength edge (> 1160 nm), the transmitted intensity
decreases rapidly. On the short wavelength side, from 900 nm to 1000 nm, the transmission
oscillates and then decreases rapidly when approaching 900 nm. This spectral transmittance is
not affected by bending the fiber, as long as the bend radius is greater than a few millimeters.
We believe that the region between 900 nm and 1000 nm lies within the bandgap of the
cladding material, but that losses in this range are significantly increased by virtue of anti-
crossings [7,8] with several interface modes, which act as an additional loss mechanism.
We have studied near-field patterns over the photonic bandgap wavelength range,
including both low-loss and interface-mode regions. We used a short length (5 m) of HC-PCF
and a fiber-based optical supercontinuum [11] as a light source. The supercontinuum was
passed through a monochromator (3 nm bandwidth) and then coupled into the HC-PCF using
an objective lens. The output end of the HC-PCF was imaged onto a digital CCD camera with
high magnification, so that the output field patterns could be studied as a function of
wavelength. Images were recorded at 1 nm intervals. We were unable to cover the long
wavelength edge of the guiding region (beyond the supercontinuum pump wavelength of 1064

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1479
0

-1

Transmitted spectrum (dB)


-2
200

-3

Attenuation (dB/km)
150

100
-4

50

-5
0
950 1000 1050 1100 1150 1200

Wavelength (nm)
-6
900 950 1000 1050 1100 1150 1200

Wavelength (nm)

Fig. 2. Spectrum of a broadband light source (tungsten lamp) transmitted through a 5m length
of the HC-PCF. The inset shows the attenuation recorded using a cutback measurement on 85
m of the same fiber.

-2
Transmitted spectrum (dB)

-4

-6

-8

-10
900 910 920 930 940 950 960 970 980 990 1000

Wavelength (nm)

Fig 3. Enlarged view of the interface-mode region of the transmission curve from Fig. 2, with
observed near-field intensity maps at the fiber output at wavelengths corresponding to some of
the intensity mimima and maxima. The entire sequence can be viewed as a movie. (avi, 505kB)

nm) because of the limited spectral response of our CCD camera. The data were corrected for
the spectral profile of the supercontinuum source and the spectral response of the CCD
camera.
An expanded view of the transmission curve through the interface-mode region is shown
in Fig. 3, along with some observed modal field patterns observed at high- and low-
transmission wavelengths. A complete data series (spanning 902 nm to 1042 nm) can be
watched as a movie by clicking on the link from the Fig. 3 caption. Over the low-loss region

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1480
at wavelengths beyond 1000 nm, the mode has a quasi-Gaussian intensity pattern. However,
moving into the bandgap from the short-wavelength side, we observe that the sharp variations
in the measured attenuation are accompanied by radical changes in the observed output modal
field patterns. Around the transmission minima, the strongly-peaked hollow-core mode is
replaced by weaker modal patterns which are less strongly peaked in the air core. Clearly, this
will have a substantial effect on the damage threshold and nonlinear response of the fibers, as
well as on the dispersion [2] and attenuation.
It is interesting to consider whether the birefringence in the fiber is sufficiently high to split
the interface-mode crossings enough to be observable spectrally. Experimentally, this is
complicated by the low birefringence of the fundamental mode in this fiber, which means that
polarization is not maintained along the 5 m fiber length. By using a polarization analyzer
between the HC-PCF and the camera we can readily observe both different modal field
patterns (shown in Fig. 4(a)) and a different spectral response as a function of output
polarization. Fig. 4a shows that at 979 nm, polarization 2 is more lossy and less peaked in the
hollow core, whereas the situation is reversed by 990 nm. However, we cannot ascribe
discrete loss peaks to specific polarization states, most probably because of poor control of
polarization over the fiber length, a relatively small spectral splitting and perhaps a lack of
coincidence of the polarization axes for the interface and air modes.
Polarization 1 Polarization 2

979 nm

Linear scale
1021 nm

982 nm

Linear scale
990 nm 1040 nm
(a) (b)
Fig. 4. (a) Polarized near-field intensity patterns recorded using low numerical aperture
excitation as a function of wavelength (log scale). (b) Near-field patterns at two different
wavelengths using a higher numerical aperture excitation with the same output polarization
(linear scale). One of these (1021 nm) shows interface-mode features, even though these are
not apparent in the fundamental mode at this wavelength.
By using a high-power (60×) objective lens, we were also able to excite higher-order
modes in the hollow core (see Fig. 4(b)). Whereas the fundamental mode consists of a single

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1481
polarization pair, the next set of modes consists of four closely-spaced modes, all generally
“donut” shaped. In our experiment we expect to excite some superposition of these. The
attenuation of these modes is higher than for the fundamental, so that they are not usually
observed in long lengths of fiber. Also, they are sufficiently well split in propagation constant
from the fundamental to not be coupled to it by bends and twists in the fiber. As these modes
have different propagation constants to the fundamental, they cross with the interface modes
at different, longer wavelengths. As an example, we present in Fig. 4(b) images of the second
core mode crossing an interface mode in a wavelength range where the fundamental mode is
not affected by interface modes.
4. Fiber core designs for beam delivery
In traditional fibers, delivery of nanosecond pulses (or CW beams) is limited by fiber damage
and nonlinear effects such as Raman generation. These limitations can be substantially
relieved by using HC-PCF. The influence of the HC-PCF design manifests in two ways: the
overlap of the guided mode with a Gaussian beam determines the beam launch efficiency
(assuming a high-quality input beam) and thus the percentage of incident light that couples to
the fundamental guided mode. Secondly, the guided mode profile sets the percentage of light
guided in the air and the guided mode intensity in the silica parts of the fiber, in turn
influencing the nonlinear response. In this section, we compare two designs of hollow core
fiber with reference to these important criteria.

Fig. 5. 7-cell and 19-cell fiber structures and examples of computed guided-mode field
patterns.

Examples of 7-cell-core and 19-cell-core fiber designs are shown in Fig. 5, along with
examples of computed guided-mode field patterns in these two structures. We use a plane-
wave expansion method to compute the different guided modes of these fibers for a fixed

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1482
cladding structure. The chosen cladding structure has an air-fraction of 91%, and
approximates the shapes of the air holes as hexagons with rounded corners. The thickness of
the silica struts forming the framework of the cladding was fixed to be 0.0304 times the pitch,
as was the core wall. The radii of the circles used to form the rounded hexagons was chosen
so that 40% of each strut was of uniform thickness. This cladding creates a band gap for the
fundamental core-guided mode extending over 24% of the central wavelength. The basic
features of the observed attenuation spectra, i.e. the width and central wavelength of the band
gap, are well reproduced by the computation. Modeling structures similar to that shown, we
consistently find that there are interface-mode crossings on the short-wavelength side of the
band gap.
In order to compare the 7-cell and 19-cell structures, the fundamental mode-field profiles
are calculated for both designs at a range of wavelengths that span the entire band gap. For
each structure, an optimum wavelength is chosen that maximizes the percentage of light
propagating in the air parts of the fiber. At these wavelengths, the modes are not affected by
anti-crossings with interface modes. Table 1 compares values for the percentage light-in-air of
the fundamental modes, the peak intensity in the silica (normalized to the maximum intensity
at the core center) and the optimized overlap with a Gaussian beam. The nonlinear phase shift
(computed as in reference [9]), calculated per MW of power and per m of length, is also
included in the table.
The results in Table 1 show the potential advantages of using a 19-cell fiber to guide
high-power lasers. The 19-cell fiber possesses a larger mode area, and since the normalized
peak intensity in silica is approximately the same for both designs, the absolute peak intensity
is lower in the 19-cell case, for a given power. Furthermore, the percentage of light traveling
in the silica parts of the fiber is reduced from 0.8% to 0.2% by moving to the larger core. This
reduced overlap with the silica combined with the increased area gives rise to a nonlinear
phase shift that is significantly lower in the 19-cell design. The nonlinear phase shift due to
the air in the core then dominates the nonlinear response (by a ratio 10 to 1), whereas in the 7-
cell structure the contributions to the phase shift of the air and silica parts of the fiber are
approximately equal. In addition to the advantage of reduced overlap with the silica, the mode
overlap with a Gaussian beam is considerably better for the 19-cell structure, giving greater
beam launch efficiency and decreased fraction of power incident upon the microstructured
cladding. 19-cell designs are thus likely to offer superior performance for transmission of CW
or ns pulsed beams, as long as the wavelength of interest is well away from an interface-mode
anti-crossing.
Further experimental and numerical studies have shown that the 19-cell design has a
higher density of interface modes than the 7-cell case. This must be expected due to the longer
core perimeter associated with the larger core. The greater number of anti-crossings between
these modes and the fundamental mode can result in a narrowing of the low-loss

Table 1. Properties of 7-cell and 19-cell fibers.

7-Cell Structure 19-Cell Structure

Percentage light-in-air 99.2% 99.8%


Peak intensity in silica (normalized 0.095 0.097
to max. intensity)

Gaussian overlap 90.4% 96.1%


Nonlinear phase shift (MW-1.m-1 ) 4.39π 1.34π

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1483
transmission windows in the band gap [2]. This is not a significant problem for transmission
of ns pulsed or CW beams, which are intrinsically spectrally narrow, but becomes important
for shorter pulse transmission.
The situation for delivery of sub-ps or fs pulses is rather different. Here, the primary
limitations arise from the bandwidth of the pulse and the dispersion of the fiber. Pulses can be
delivered either as high-energy solitons [4,9] or by pre-chirping the pulses (either with or
without further spectral broadening) and then recompression in a length of hollow-core fiber
[12]. In the first case, one challenge is to adjust the nonlinear phase shift to enable efficient
soliton excitation with the available pulse energy, while maintaining both a low dispersion
and dispersion slope. Fibers with 7-cell cores have been shown to be capable of supporting
relatively long solitons (several hundred fs) when excited using amplified modelocked laser
systems [4,9]. This has not yet been demonstrated using unamplified systems because the
nonlinear phase shift is usually too low, even in a 7-cell defect fiber. However, the nonlinear
response could be increased by using a lower air-filling-fraction cladding, a thicker core
surround or by using a smaller core. A second challenge, which is relevant for linear (pre-
chirped) propagation as well, is to form fibers which have the required level of second-order
dispersion (GVD) but have suitably low higher-order dispersion over the bandwidth of the
laser. We would expect that a larger core size – i.e. a 19-cell defect – would result in lower
waveguide dispersion and hence lower dispersion slope. However, the increased density of
interface modes associated with a larger core invalidates that argument, breaking up the
transmission band into a number of low-loss windows within which the dispersion is
adversely affected [8,10]. Consequently, we believe that 7-cell-core fibers are presently more
suited for the delivery of sub-picosecond or femtosecond pulses.
This conclusion may change in the future. Numerical computations show that the number
and position of interface modes in a 19-cell structure are very sensitive to deformations of the
core and cladding of the fiber. Improvements to the manufacture of 19-cell designs may
broaden the low-loss and low-dispersion transmission windows to the extent that they become
suitable for shorter pulses, allowing all the advantages of using a larger core to be realized.
5. Conclusions
As for conventional fibers, HC-PCF designs with larger cores are more suitable for
transmission of high-power or high-energy laser beams, because of their increased damage
threshold and reduced nonlinear response. However, the low-loss transmission bandwidth in
current 19-cell fibers is significantly reduced as a result of the greater density of interface
modes and their associated anti-crossings with the fundamental guided mode. We have
observed the effects of interface-mode anti-crossings on the fundamental and higher-order
core modes in a 7-cell fiber, and shown that the ideal properties of the guided modes are
substantially destroyed in the vicinity of such a crossing. Future improvements in the
manufacture of the 19-cell design may broaden the low-loss transmission windows to a level
suitable for sub-ps and fs pulse delivery, but at this stage, the smaller-core fibers offer a lower
interface mode density, giving fewer mode anti-crossings, more favourable dispersion
characteristics and a broader bandwidth for ultra-short pulse delivery. Both 7-cell and 19-cell
fiber designs will find applications in the delivery of laser beams in the future, depending on
the specific application being developed.

#3973 - $15.00 US Received 3 March 2004; revised 6 April 2004; accepted 6 April 2004
(C) 2004 OSA 19 April 2004 / Vol. 12, No. 8 / OPTICS EXPRESS 1484

Das könnte Ihnen auch gefallen