Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to main content

On Plenoptic Multiplexing and Reconstruction

  • Published:
International Journal of Computer Vision Aims and scope Submit manuscript

Abstract

Photography has been striving to capture an ever increasing amount of visual information in a single image. Digital sensors, however, are limited to recording a small subset of the desired information at each pixel. A common approach to overcoming the limitations of sensing hardware is the optical multiplexing of high-dimensional data into a photograph. While this is a well-studied topic for imaging with color filter arrays, we develop a mathematical framework that generalizes multiplexed imaging to all dimensions of the plenoptic function. This framework unifies a wide variety of existing approaches to analyze and reconstruct multiplexed data in either the spatial or the frequency domain. We demonstrate many practical applications of our framework including high-quality light field reconstruction, the first comparative noise analysis of light field attenuation masks, and an analysis of aliasing in multiplexing applications.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Explore related subjects

Discover the latest articles, news and stories from top researchers in related subjects.

Notes

  1. For a 2D sensor image the bases are periodic in both spatial dimensions, but we will omit the second one in our notation of \(k\) for clarity.

  2. Following Sect. 3.4, we omit periodicity in the second spatial dimension in our notation of \(t\) for clarity.

  3. For other sampling patterns, corresponding, e.g., hexagonal, masking functions would be used. In addition, apodization functions can be used to reduce ringing artifacts in the reconstruction at the expense of decreasing the effective resolution, see e.g., (Veeraraghavan et al. 2007). This does not affect the proof.

References

  • Adelson, E.,& Wang, J. (1992). Single lens stereo with a plenoptic camera. IEEE Transactions on Pattern Analysis and Machine, 14(2), 99–106.

    Article  Google Scholar 

  • Adelson, E. H.,& Bergen, J. R. (1991). The plenoptic function and the elements of early vision. In Computational models of visual processing (pp. 3–20). Cambridge, MA: MIT Press.

  • Aggarwal, M.,& Ahuja, N. (2004). Split aperture imaging for high dynamic range. International Journal of Computer Vision, 58(1), 7–17.

    Google Scholar 

  • Agrawal, A., Gupta, M., Veeraraghavan, A.,& Narasimhan, S. (2010a). Optimal coded sampling for temporal super-resolution (pp. 374–380). In Proceedings IEEE CVPR.

  • Agrawal, A., Veeraraghavan, A.,& Raskar, R. (2010b). Reinterpretable imager: Towards variable post-capture space, angle and time resolution in photography. In: Proceedings of eurographics (pp. 1–10).

  • Alleyson, D., Süsstrunk, S.,& Hérault, J. (2005). Linear demosaicing inspired by the human visual system. IEEE Transactions On Instrumentation and Measurement, 14(4), 439–449.

    Google Scholar 

  • Bando, Y., Chen, B . Y.,& Nishita, T. (2008). Extracting depth and matte using a color-filtered aperture. ACM Transactions on Graphics Siggraph Asia, 27(5), 134.

    Google Scholar 

  • Bayer, B. E. (1976). Color imaging array. US Patent, 3,971,065.

  • Ben-Ezra, M., Zomet, A.,& Nayar, S. (2005). Video superresolution using controlled subpixel detector shifts. IEEE Transactions on Pattern Analysis and Machine Intelligence, 27(6), 977–987.

    Article  Google Scholar 

  • Bishop, T., Zanetti, S.,& Favaro, P. (2009). Light-field superresolution. In Proceedings of ICCP (pp. 1–9 ).

  • Bub, G., Tecza, M., Helmes, M., Lee, P.,& Kohl, P. (2010). Temporal pixel multiplexing for simultaneous high-speed, high-resolution imaging. Nature Methods, 7, 209–211.

    Article  Google Scholar 

  • Compton, J. (2007). Color filter array 2.0. http://johncompton.pluggedin.kodak.com

  • Debevec, P. E.,& Malik, J. (1997). Recovering high dynamic range radiance maps from photographs. In Proceedings of ACM siggraph (pp. 369–378).

  • Georgiev, T.,& Lumsdaine, A. (2010). Rich imamge capture with plenoptic cameras. In Proceedings of ICCP (pp. 1–8).

  • Georgiev, T., Intwala, C., Babacan, S.,& Lumsdaine, A. (2008). Unified frequency domain analysis of lightfield cameras (pp. 224–237). In Proceedings of ECCV.

  • Gortler, S., Grzeszczuk, R., Szelinski, R.,& Cohen, M. (1996). The lumigraph. In: Proceedings of ACM siggraph (pp. 43–54).

  • Gottesman, S . R.,& Fenimore, E . E. (1989). New family of binary arrays for coded aperture imaging. Applied Optics, 28(20), 4344–4352.

    Article  Google Scholar 

  • Greivenkamp, J. (1990). Color dependant optical prefilter for the suppression of aliasing artifacts. Applied Optics, 29(5), 676–684.

    Article  Google Scholar 

  • Gupta, M., Agrawal, A., Veeraraghavan, A.,& Narasimhan, S. G. (2010). Flexible voxels for motion-aware videography. In Proceedings of ECCV (pp. 100–114).

  • Horstmeyer, R., Euliss, G., Athale, R.,& Levoy, M. (2009). Flexible multimodal camera using a light field architecture. In Proceedings of ICCP (pp. 1–8).

  • Ives, F. E. (1903). Parallax stereogram and process of making same. U.S. Patent 725,567.

  • Lanman, D., Raskar, R., Agrawal, A.,& Taubin, G. (2008). Shield fields: Modeling and capturing 3D occluders. ACM Transactions on Graphics (Siggraph Asia), 27(5), 131.

    Google Scholar 

  • Levin, A.,& Durand, F. (2010). Linear view synthesis using a dimensionality gap light field prior. In Proceedings of IEEE CVPR (pp. 1–8).

  • Levin, A., Freeman, W. T.,& Durand, F. (2008). Understanding camera trade-offs through a bayesian analysis of light field projections. In Proceedings of ECCV (pp. 88–101).

  • Levin, A., Hasinoff, S. W., Green, P., Durand, F.,& Freeman, W. T. (2009). 4D frequency analysis of computational cameras for depth of field extension. ACM Transactions on Graphics (Siggraph), 28(3), 97.

    Article  Google Scholar 

  • Levoy, M.,& Hanrahan, P. (1996). Light field rendering. In Proceedings of ACM siggraph (pp. 31–42).

  • Levoy, M., Ng, R., Adams, A., Footer, M.,& Horowitz, M. (2006). Light field microscopy. ACM Transactions on Graphics (Siggraph), 25(3), 924–934.

    Article  Google Scholar 

  • Li, X., Gunturk, B.,& Zhang, L. (2008). Image demosaicing: A systematic survey (pp 68221J–68,221J–15). In SPIE Conference on visual communication and image processing.

  • Lippmann, G. (1908). La Photographie Intégrale. Academie des Sciences, 146, 446–451.

    Google Scholar 

  • Lumsdaine, A.,& Georgiev, T. (2009). The focused plenoptic camera. In Proceedings of ICCP (pp. 1–8).

  • McGuire, M., Matusik, W., Pfister, H., Chen, B., Hughes, J. F.,& Nayar, S. K. (2007). Optical splitting trees for high-precision monocular imaging. IEEE Transaction on Knowledge& Data Engineering, 27(2), 32–42.

    Google Scholar 

  • Mitsunaga, T.,& Nayar, S. K. (1999). Radiometric self calibration. In Proceedings of IEEE CVPR (pp. 374–380).

  • Narasimhan, S.,& Nayar, S. (2005). Enhancing resolution along multiple imaging dimensions using assorted pixels. IEEE Transactions on Pattern Analysis and Machine Intelligence, 27(4), 518–530.

    Article  Google Scholar 

  • Ng, R. (2005). Fourier slice photography. ACM Transactions on Graphics (Siggraph), 24(3), 735–744.

    Article  Google Scholar 

  • Reddy, D., Veeraraghavan, A.,& Chellappa, R. (2011). P2C2: Programmable pixel compressive camera for High speed imaging. In Proceedings of IEEE CVPR (pp. 1–8).

  • Sajadi, B., Majumder, A., Hiwada, K., Maki, A.,& Raskar, R. (2011). Switchable primaries using shiftable layers of color filter arrays. ACM Transactions on Graphics (Siggraph), 30(3), 1–8.

    Google Scholar 

  • Schechner, Y.,& Nayar, S. (2003). Generalized mosaicing: high dynamic range in a wide field of view. International Journal of Computer Vision, 53(3), 245–267.

    Google Scholar 

  • Schechner, Y., Nayar, S.,& Belhumeur, P. (2007). Multiplexing for optimal lighting. IEEE Transactions on Pattern Analysis and Machine Intelligence, 29(8), 1339–1354.

    Article  Google Scholar 

  • Takhar, D., Laska, J. N., Wkin, M., Duarte, M. F., Baron, D., Sarvotham, S., Kelly, K. F.,& Baraniuk, R. G. (2006). A new compressive imaging camera architecture using optical-domain compression. IS&T/SPIE computational imaging IV 6065.

  • Veeraraghavan, A., Raskar, R., Agrawal, A., Mohan, A.,& Tumblin, J. (2007). Dappled photography: mask enhanced cameras for heterodyned light fields and coded aperture refocussing. ACM Transactions on Graphics (Siggraph), 26(3), 69.

    Article  Google Scholar 

  • Veeraraghavan, A., Raskar, R., Agrawal, A., Chellappa, R., Mohan, A.,& Tumblin, J. (2008). Non-refractive modulators for encoding and capturing scene appearance and depth. In Proceedings of IEEE CVPR (pp. 1–8).

  • Veeraraghavan, A., Reddy, D.,& Raskar, R. (2011). Coded strobing photography: compressive sensing of high speed periodic videos. IEEE Transactions on Pattern Analysis and Machine Intelligence, 33(4), 671–686.

    Article  Google Scholar 

  • Wang, S.,& Heidrich, W. (2004). The design of an inexpensive very high resolution scan camera system. Computer Graphics Forum (Eurographics), 23(10), 441–450.

    Article  Google Scholar 

  • Wenger, A., Gardner, A., Tchou, C., Unger, J., Hawkins, T.,& Debevec, P. (2005). Performance relighting and reflectance transformation with time-multiplexed illumination. ACM Transactions on Graphics (Siggraph), 24(3), 756–764.

    Article  Google Scholar 

  • Wetzstein, G., Ihrke, I.,& Heidrich, W. (2010). Sensor saturation in Fourier multiplexed imaging. In Proceedings of IEEE CVPR (pp. 1–8).

  • Wetzstein, G., Ihrke, I., Lanman, D.,& Heidrich, W. (2011). State of the art in computational plenoptic imaging. In Proceedings of eurographics (STAR) (pp 1–24).

  • Wilburn, B., Joshi, N., Vaish, V., Talvala, E . V., Antunez, E., Barth, A., et al. (2005). High performance imaging using large camera arrays. ACM Transactions on Graphics (Siggraph), 24(3), 765–776.

    Article  Google Scholar 

  • Wuttig, A. (2005). Optimal transformations for optical multiplex measurements in the presence of photon noise. Applied Optics, 44(14), 2710–2719.

    Article  Google Scholar 

  • Yasuma, F., Mitsunaga, T., Iso, D.,& Nayar, S. K. (2010). Generalized assorted pixel camera: post-capture control of resolution.In IEEE transaction Im proceedings dynamic range and spectrum (p. 99).

Download references

Acknowledgments

We thank Dolby Canada for their support and the anonymous reviewers for their insightful feedback. Gordon Wetzstein was supported by a UBC Four Year Fellowship, an NSERC Postdoctoral Fellowship, and the DARPA SCENICC program. Wolfgang Heidrich was supported under the Dolby Research Chair in Computer Science at UBC. Ivo Ihrke was supported by a Feodor-Lynen Fellowship of the Humboldt Foundation, Germany.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gordon Wetzstein.

Appendices

Appendix A: Proof of PSM Theorem

Throughout this paper, we assume that the spatial basis functions \(\varvec{\sigma }_j ( \vec {x})\) are super-pixel-periodicFootnote 2, i.e., \( \varvec{\sigma }_j ( \vec {x}) = \varvec{\sigma }_j ( \vec {x}+ t \Delta \vec {x}_s), \, \forall t \in \mathbb{Z }.\) The offset between successive super-pixels is denoted as \(\Delta \vec {x}_s.\) This notation allows us to define a sampling operator as

$$\begin{aligned} \mathrm{I\!I\!I}_k(\vec {x}) = \sum \limits _{t \in \mathbb{Z }} \delta \left( \vec {x}+ t \Delta \vec {x}_s+ \Delta \vec {x}_k \right), \end{aligned}$$
(23)

where \(\Delta \vec {x}_k\) is the offset of individual samples within each super-pixel. This sampling operator basically extracts a sub-image or channel \(k=1 \ldots M\) from an interleaved sensor image, where only corresponding sub-pixels within the super-pixels are included in each channel. Sampling such a channel \(\widetilde{\varvec{i}}_k (\vec {x})\) from a sensor image \(\varvec{i}(\vec {x}),\) in combination with Eq. 4, results in the following expression

$$\begin{aligned} \widetilde{\varvec{i}}_k (\vec {x})&= \mathrm{I\!I\!I}_k(\vec {x}) \varvec{i}(\vec {x}) = \mathrm{I\!I\!I}_k(\vec {x}) \left( \sum \limits _{j=1}^{N} \varvec{\sigma }_j (\vec {x}) \varvec{\rho }_j \left( \vec {x}\right)\right) \nonumber \\&= \sum \limits _{j=1}^{N} \varvec{\sigma }_j^k \mathrm{I\!I\!I}_k(\vec {x}) \varvec{\rho }_j \left( \vec {x}\right). \end{aligned}$$
(24)

As discussed in Sect. 3.3, the spatial basis functions \(\varvec{\sigma }_j (\vec {x})\) become spatially-invariant constants \(\varvec{\sigma }_j^k\) in the sampled channels because of spatial periodicity of the basis. Each channel \(\tilde{\varvec{i}}_k (\vec {x})\) is thus associated with \(N\) constants \(\varvec{\sigma }_j^k\) defined by the spatial basis. Reconstructing a channel \(\varvec{i}_k (\vec {x})\) from its sampled representation \(\tilde{\varvec{i}}_k (\vec {x})\) is performed by convolving with a reconstruction filter kernel \(\varvec{f}(\vec {x}):\)

$$\begin{aligned} \varvec{i}_k (\vec {x}) = \widetilde{\varvec{i}}_k (\vec {x}) \otimes \varvec{f}(\vec {x}). \end{aligned}$$
(25)

According to the sampling theorem, the original signal must be spatially band-limited for this reconstruction to be a faithful representation.

Due to the spatial basis being a set of constants at every position in the interpolated channels \(\varvec{i}_k (\vec {x}),\) and the plenoptic basis being spatially-invariant, a convolution with the filter kernel can be formulated as

$$\begin{aligned} \varvec{i}_k (\vec {x})&= \left( \sum \limits _{j=1}^{N} \varvec{\sigma }_j^k \mathrm{I\!I\!I}_k(\vec {x}) \varvec{\rho }_j \left( \vec {x}\right)\right) \otimes \varvec{f}(\vec {x})\\&= \sum \limits _{j=1}^{N} \varvec{\sigma }_j^k \int _\fancyscript{P}\varvec{\pi }_j (\vec {p}) \bigg ( \mathrm{I\!I\!I}_k(\vec {x}) \varvec{l}_\lambda (\vec {x},\vec {p})\otimes \varvec{f}(\vec {x})\bigg ) d\vec {p}\nonumber \end{aligned}$$
(26)

Equation 26 shows that all channels \(\varvec{i}_k (\vec {x})\) are locally related to the sampled and reconstructed plenoptic function via a linear combination. It also shows that applying a linear filter to the measured channels, i.e., before decorrelation, is equivalent to applying the same filter to the plenoptic function itself, i.e., after decorrelation. \(\square \)

Appendix B: Proof of PFM Theorem

The proof of Theorem 2 follows Equations 6-8. To provide additional detail, we start with Eq. 8

$$\begin{aligned} \mathcal F _{x} \left\{ \varvec{i}(\vec {x}) \right\} \! = \! \sum \limits _{k=1}^M \left( \! \delta \left( {\vec {\omega }_x- \Delta \vec {\omega }_x^k} \right) \! \otimes \! \sum \limits _{j=1}^N \hat{\varvec{\sigma }}_j^{\,k}\hat{\varvec{\rho }}_j\left(\vec {\omega }_x\right) \! \right), \end{aligned}$$
(27)

where \(k \Delta \vec {\omega }_x\) has been replaced by \(\Delta \vec {\omega }_x^k,\) a vector to the center frequency of a Fourier tile (Fig. 13, right), to allow for generalized sampling patterns. If the signal is properly band-limited the image’s Fourier transform separates into disjoint sets, each encoding one Fourier tile \(\hat{\varvec{i}}_k (\vec {\omega }^{\prime }_x)\) (Fig. 13, right). To separate the notation for spatial frequencies in the sensor image and for Fourier tiles we use the substitution \(\vec {\omega }^{\prime }_x= \vec {\omega }_x- \Delta \vec {\omega }_x^k\).

Fig. 13
figure 13

Information overlaps in the Fourier domain if the signal is not suitably band-limited (left). With the appropriate band-limitation, the Fourier representation decomposes into distinct correlated Fourier tiles \(\hat{\varvec{i}}_k.\) The arrowin the right figure indicates the position \(\Delta \vec {\omega }_x^k\) of the Dirac peaks, i.e., the center frequencies, of a Fourier tile

The Fourier tiles can be cropped from the Fourier transformed sensor image by applying a rect filter:Footnote 3

$$\begin{aligned} \hat{\varvec{i}}_k (\vec {\omega }^{\prime }_x)&= \mathrm rect ^k(\vec {\omega }^{\prime }_x) \delta \left( {\vec {\omega }^{\prime }_x} \right) \otimes \sum \limits _{j=1}^N \hat{\varvec{\sigma }}_j^{\,k}\hat{\varvec{\rho }}_j\left(\vec {\omega }^{\prime }_x\right) \\&= \mathrm rect ^k(\vec {\omega }^{\prime }_x) \sum \limits _{j=1}^N \hat{\varvec{\sigma }}_j^{\,k}\int _\fancyscript{P}\varvec{\pi }(\vec {p}) \hat{\varvec{l}}_\lambda (\vec {\omega }^{\prime }_x, \vec {p}) d\vec {p}.\nonumber \end{aligned}$$
(28)

The convolution with a Dirac train in Eq. 8 reduces to a convolution with a single Dirac peak \(\delta \left( {\vec {\omega }^{\prime }_x} \right)\) because of band-limitation. This is a unit operation and thus removed from the equations. \(\square \)

In addition, by inverse Fourier transforming and sampling the Fourier tile \(\hat{\varvec{i}}_k (\vec {\omega }^{\prime }_x)\) we see that the reconstruction filter is indeed a sinc:

$$\begin{aligned} \varvec{i}_k(\vec {x}) \! = \! \sum \limits _{j=1}^{N} \varvec{\sigma }_j^k \int _\fancyscript{P}\varvec{\pi }_j (\vec {p}) \bigg ( \mathrm{I\!I\!I}_k(\vec {x}) \varvec{l}_\lambda (\vec {x},\vec {p})\otimes \mathrm sinc (\vec {x}) \!\bigg ) d\vec {p}\nonumber \\ \end{aligned}$$
(29)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wetzstein, G., Ihrke, I. & Heidrich, W. On Plenoptic Multiplexing and Reconstruction. Int J Comput Vis 101, 384–400 (2013). https://doi.org/10.1007/s11263-012-0585-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11263-012-0585-9

Keywords