simultaneous, accurate measurement of the 3d position and ......simultaneous, accurate measurement...

16
Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlund a,1 , Matthew D. Lew a,b,1 , Adam S. Backer a,c , Steffen J. Sahl a , Ginni Grover d , Anurag Agrawal d , Rafael Piestun d , and W. E. Moerner a,2 Departments of a Chemistry and b Electrical Engineering and c Institute for Computational and Mathematical Engineering, Stanford University, Stanford, CA 94305; and d Department of Electrical, Computer, and Energy Engineering, University of Colorado, Boulder, CO 80309 Contributed by W. E. Moerner, September 25, 2012 (sent for review July 24, 2012) Recently, single molecule-based superresolution uorescence mi- croscopy has surpassed the diffraction limit to improve resolution to the order of 20 nm or better. These methods typically use image tting that assumes an isotropic emission pattern from the single emitters as well as control of the emitter concentration. However, anisotropic single-molecule emission patterns arise from the transi- tion dipole when it is rotationally immobile, depending highly on the molecules 3D orientation and z position. Failure to account for this fact can lead to signicant lateral (x, y) mislocalizations (up to 50200 nm). This systematic error can cause distortions in the reconstructed images, which can translate into degraded resolu- tion. Using parameters uniquely inherent in the double-lobed na- ture of the Double-Helix Point Spread Function, we account for such mislocalizations and simultaneously measure 3D molecular orien- tation and 3D position. Mislocalizations during an axial scan of a single molecule manifest themselves as an apparent lateral shift in its position, which causes the standard deviation (SD) of its lat- eral position to appear larger than the SD expected from photon shot noise. By correcting each localization based on an estimated orientation, we are able to improve SDs in lateral localization from 2× worse than photon-limited precision (48 vs. 25 nm) to within 5 nm of photon-limited precision. Furthermore, by averaging many estimations of orientation over different depths, we are able to improve from a lateral SD of 116 (4× worse than the photon- limited precision; 28 nm) to 34 nm (within 6 nm of the photon limit). T he recent emergence of superresolution far-eld optical mi- croscopy techniques has provided a means for attaining res- olution beyond the diffraction limit (250 nm) in noninvasive uorescence imaging of biological structures (1, 2). Some of these techniques [including (f)PALM (3, 4), STORM (5), and PAINT (6)] rely on precise localization of sparse subsets of single-molecule (SM) emitters to surpass the diffraction limit by up to an order of magnitude (precisions of tens of nanometers). Collectively, these SM-based superresolution techniques can be grouped under the name SM Active Control Microscopy (SMACM), because they all rely on using various experimental strategies (photoactivation, switching, blinking additives, etc.) to maintain a very low concentration of emitters in each imaging frame, enabling the localization of SMs without overlap. Typi- cally, the SM tting uses estimators that assume isotropic emis- sion, i.e., that the center of the photon distribution of an SM image corresponds directly to the true position of the molecule. Examples of these estimators include centroid nding, least- squares tting to a 2D Gaussian function, and maximum likeli- hood methods that assume isotropic emitters. However, immobile uorescing SMs produce an inherently anisotropic emission pattern that depends on the orientation of the SM emission dipole moment relative to the optical axis (7, 8). The work by Enderlein et al. (9) has shown that tting such an SM image to a 2D Gaussian can result in position errors of tens of nanometers for molecules located in the microscopes focal plane. Even more strikingly, the work by Engelhardt et al. (10) noted that, with modest defocusing (z 300 nm), the position error associated with tting to a centroid can exceed 100 nm for certain SM dipole orientations. If labels are sufciently rota- tionally mobile such that they explore much of the orientation space within a single acquisition, this effect is averaged away, and accuracy can be recovered. However, in some cases, labels of biological structures can exhibit well-dened orientations (11). Furthermore, uorophores can be purposely anchored to convey orientation information about biological macromolecules, such as work in various SM studies on motor protein translocation (1214). Although this position error has important implications for 2D SMACM techniques, the implications for 3D SMACM techniques are even more signicant. Namely, techniques such as astigmatism (15), multiplane (16), iPALM (17), and Double- Helix Point Spread Function (DH-PSF) (18) imaging depend explicitly on precise 3D localization of SMs over an extended depth of eld of up to 2 μm. Failure to account for this dipole orientation effect can clearly lead to large position inaccuracies that severely limit the superresolving capabilities of these tech- niques. Of these 3D methods, the DH-PSF is uniquely suited to address orientation effects, because its double-lobed shape results from the superposition of various waves in the micro- scopes pupil plane that converge and interfere in the image plane; the intensities of these waves are strongly affected by the dipole radiation pattern of SMs. There are many established methods for determining dipole orientation of single uorophores. Approaches have been developed that rely on excitation and/or emission with multiple polarizations (19, 20), introduction of defocus and pattern match- ing (21), direct imaging of pupil functions (22), and use of an- nular illumination to create characteristic eld distributions (23) to name a few. The alternating measurement of 2D position and orientation has also been addressed (24). Two groups considered simultaneous 2D localization and orientation tting for mole- cules located in the focal plane (25) or molecules at a known defocus (26). However, neither explicitly addresses the correc- tion of systematic errors on the order of 50100 nm, and these methods require ne sampling and accurate tting of detailed patterns in SM images. This paper shows simultaneous measure- ment of precise and accurate 3D localization and molecular orientation. We account for and correct large localization errors using an adaptation of the established DH-PSF method. The basics of the DH-PSF microscope have been described in detail elsewhere (18). Briey, the DH-PSF causes a single uo- rescent emitter to appear on the detector as two closely spaced lobes. Here, we use an estimator based on tting the two lobes with two Gaussian functions. Precise x and y localization can be extracted from the midpoint (x, y) position between the two lobes, whereas precise z localization is determined by the angle of the axis connecting the two lobes. As an emitter is moved in z, the DH-PSF revolves, effectively tracing out a double helix along Author contributions: M.P.B., M.D.L., R.P., and W.E.M. designed research; M.P.B., M.D.L., A.S.B., and G.G. performed research; S.J.S. and A.A. contributed new reagents/analytic tools; M.P.B., M.D.L., and A.S.B. analyzed data; and M.P.B., M.D.L., A.S.B., A.A., R.P., and W.E.M. wrote the paper. The authors declare no conict of interest. 1 M.P.B. and M.D.L. contributed equally to this work. 2 To whom correspondence should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1216687109/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1216687109 PNAS | November 20, 2012 | vol. 109 | no. 47 | 1908719092 APPLIED PHYSICAL SCIENCES

Upload: others

Post on 06-Jul-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

Simultaneous, accurate measurement of the 3Dposition and orientation of single moleculesMikael P. Backlunda,1, Matthew D. Lewa,b,1, Adam S. Backera,c, Steffen J. Sahla, Ginni Groverd, Anurag Agrawald,Rafael Piestund, and W. E. Moernera,2

Departments of aChemistry and bElectrical Engineering and cInstitute for Computational and Mathematical Engineering, Stanford University, Stanford,CA 94305; and dDepartment of Electrical, Computer, and Energy Engineering, University of Colorado, Boulder, CO 80309

Contributed by W. E. Moerner, September 25, 2012 (sent for review July 24, 2012)

Recently, single molecule-based superresolution fluorescence mi-croscopy has surpassed the diffraction limit to improve resolutionto the order of 20 nm or better. These methods typically use imagefitting that assumes an isotropic emission pattern from the singleemitters as well as control of the emitter concentration. However,anisotropic single-molecule emission patterns arise from the transi-tion dipole when it is rotationally immobile, depending highly onthe molecule’s 3D orientation and z position. Failure to account forthis fact can lead to significant lateral (x, y) mislocalizations (upto ∼50–200 nm). This systematic error can cause distortions in thereconstructed images, which can translate into degraded resolu-tion. Using parameters uniquely inherent in the double-lobed na-ture of the Double-Helix Point Spread Function, we account for suchmislocalizations and simultaneously measure 3D molecular orien-tation and 3D position. Mislocalizations during an axial scan ofa single molecule manifest themselves as an apparent lateral shiftin its position, which causes the standard deviation (SD) of its lat-eral position to appear larger than the SD expected from photonshot noise. By correcting each localization based on an estimatedorientation, we are able to improve SDs in lateral localization from∼2× worse than photon-limited precision (48 vs. 25 nm) to within 5nm of photon-limited precision. Furthermore, by averaging manyestimations of orientation over different depths, we are able toimprove from a lateral SD of 116 (∼4× worse than the photon-limited precision; 28 nm) to 34 nm (within 6 nm of the photon limit).

The recent emergence of superresolution far-field optical mi-croscopy techniques has provided a means for attaining res-

olution beyond the diffraction limit (∼250 nm) in noninvasivefluorescence imaging of biological structures (1, 2). Some ofthese techniques [including (f)PALM (3, 4), STORM (5), andPAINT (6)] rely on precise localization of sparse subsets ofsingle-molecule (SM) emitters to surpass the diffraction limit byup to an order of magnitude (precisions of tens of nanometers).Collectively, these SM-based superresolution techniques canbe grouped under the name SM Active Control Microscopy(SMACM), because they all rely on using various experimentalstrategies (photoactivation, switching, blinking additives, etc.) tomaintain a very low concentration of emitters in each imagingframe, enabling the localization of SMs without overlap. Typi-cally, the SM fitting uses estimators that assume isotropic emis-sion, i.e., that the center of the photon distribution of an SMimage corresponds directly to the true position of the molecule.Examples of these estimators include centroid finding, least-squares fitting to a 2D Gaussian function, and maximum likeli-hood methods that assume isotropic emitters.However, immobile fluorescing SMs produce an inherently

anisotropic emission pattern that depends on the orientation ofthe SM emission dipole moment relative to the optical axis (7, 8).The work by Enderlein et al. (9) has shown that fitting such anSM image to a 2D Gaussian can result in position errors of tensof nanometers for molecules located in the microscope’s focalplane. Even more strikingly, the work by Engelhardt et al. (10)noted that, with modest defocusing (z = ±300 nm), the positionerror associated with fitting to a centroid can exceed 100 nm forcertain SM dipole orientations. If labels are sufficiently rota-tionally mobile such that they explore much of the orientation

space within a single acquisition, this effect is averaged away, andaccuracy can be recovered. However, in some cases, labels ofbiological structures can exhibit well-defined orientations (11).Furthermore, fluorophores can be purposely anchored to conveyorientation information about biological macromolecules, suchas work in various SM studies on motor protein translocation(12–14). Although this position error has important implicationsfor 2D SMACM techniques, the implications for 3D SMACMtechniques are even more significant. Namely, techniques such asastigmatism (15), multiplane (16), iPALM (17), and Double-Helix Point Spread Function (DH-PSF) (18) imaging dependexplicitly on precise 3D localization of SMs over an extendeddepth of field of up to ∼2 μm. Failure to account for this dipoleorientation effect can clearly lead to large position inaccuraciesthat severely limit the superresolving capabilities of these tech-niques. Of these 3D methods, the DH-PSF is uniquely suited toaddress orientation effects, because its double-lobed shaperesults from the superposition of various waves in the micro-scope’s pupil plane that converge and interfere in the image plane;the intensities of these waves are strongly affected by the dipoleradiation pattern of SMs.There are many established methods for determining dipole

orientation of single fluorophores. Approaches have beendeveloped that rely on excitation and/or emission with multiplepolarizations (19, 20), introduction of defocus and pattern match-ing (21), direct imaging of pupil functions (22), and use of an-nular illumination to create characteristic field distributions (23)to name a few. The alternating measurement of 2D position andorientation has also been addressed (24). Two groups consideredsimultaneous 2D localization and orientation fitting for mole-cules located in the focal plane (25) or molecules at a knowndefocus (26). However, neither explicitly addresses the correc-tion of systematic errors on the order of 50–100 nm, and thesemethods require fine sampling and accurate fitting of detailedpatterns in SM images. This paper shows simultaneous measure-ment of precise and accurate 3D localization and molecularorientation. We account for and correct large localization errorsusing an adaptation of the established DH-PSF method.The basics of the DH-PSF microscope have been described in

detail elsewhere (18). Briefly, the DH-PSF causes a single fluo-rescent emitter to appear on the detector as two closely spacedlobes. Here, we use an estimator based on fitting the two lobeswith two Gaussian functions. Precise x and y localization can beextracted from the midpoint (x, y) position between the twolobes, whereas precise z localization is determined by the angleof the axis connecting the two lobes. As an emitter is moved in z,the DH-PSF revolves, effectively tracing out a double helix along

Author contributions: M.P.B., M.D.L., R.P., and W.E.M. designed research; M.P.B., M.D.L.,A.S.B., and G.G. performed research; S.J.S. and A.A. contributed new reagents/analytictools; M.P.B., M.D.L., and A.S.B. analyzed data; and M.P.B., M.D.L., A.S.B., A.A., R.P., andW.E.M. wrote the paper.

The authors declare no conflict of interest.1M.P.B. and M.D.L. contributed equally to this work.2To whom correspondence should be addressed. E-mail: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1216687109/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1216687109 PNAS | November 20, 2012 | vol. 109 | no. 47 | 19087–19092

APP

LIED

PHYS

ICAL

SCIENCE

S

Page 2: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

the optical axis and thus, encoding the z position in the angle ofrevolution calibrated separately by z translation of a fluorescentbead. In the experimental implementation (Fig. 1A), the DH-PSF response is generated by convolution with the standard SMimage using an appropriate phase mask at the Fourier plane ofa 4f imaging system built directly after the intermediate imageplane of a standard microscope. Typically, the phase mask isloaded onto a phase-only reflective liquid crystal spatial lightmodulator (SLM) (27–29). This type of SLM can only modulatevertically polarized light, and therefore, the emission must bepolarized before being detected. Previously, a dual-polarizationDH-PSF microscope was described, in which two orthogonalpolarizations were split by a polarizing beam splitter, the hori-zontally polarized channel was rotated with a λ/2 wave plate, andeach polarization channel was then reflected separately off theSLM at necessarily disparate angles of incidence (30). Thisscheme is not ideal for our application, because the differentangles of incidence on the SLM cause each channel to incurdissimilar aberrations. For the measurement in this paper, it wasnecessary to detect the two polarization channels with maximalchannel symmetry provided by the setup shown in Fig. 1B. Thetwo polarization channels are forced to have the same angle ofincidence on the SLM by making use of a square pyramidalmirror to deflect the beams out of the plane into the w direction(marked in Fig. 1 B–D) and onto the SLM mounted from abovewith its face to the mirror (Fig. 1 C and D). Polarized images aremeasured with two orthogonal orientations of the phase mask(Fig. 1B Inset) for reasons described below.It is well-known that splitting emission into orthogonal po-

larization channels alone yields some information about anemitter’s azimuthal (ϕ) (Fig. 1A Insets) orientation based on thecomputed linear dichroism,

LD =NT −NR

NT +NR≈ cosð2ϕÞ; [1]

in which NT and NR are the numbers of photons detected abovebackground in the transmitted and reflected polarization chan-nels, respectively, and the last equality is exact only if the par-tially depolarizing effect of high numerical aperture (N.A.)optics is ignored. Here, transmitted and reflected are defined

relative to the polarizing beam splitter (Fig. 1B). Clearly, LD isrelated to the projection of an SM dipole onto the detectionpolarizations, which in turn, is related to ϕ, but there exist de-generacies if LD is the only recorded measurement (12). Tobreak degeneracies and measure polar orientation (θ) (Fig. 1AInsets), another parameter must be measured. Interestingly, theDH-PSF uniquely offers such a parameter. Namely, the relativeintensity of the two lobes of the DH-PSF is actually a function of(z, θ, ϕ) of an SM emitter. Whereas an isotropic point sourceyields lobes of equal intensity for all z when convolved with theDH-PSF, our simulations (vide infra) show that images of SMdipoles can exhibit large lobe asymmetries (LA) for certain ori-entations at various values of defocus (Fig. 2A). Qualitatively,this asymmetry is introduced because the asymmetric pupil func-tions of SM dipoles (22) are multiplied by the DH-PSF phasemask, causing various spatial frequencies of the ordinary DH-PSF to be attenuated as a function of orientation and defocus.We quantify the lobe asymmetry as

LA =AL1 −AL2

AL1 +AL2; [2]

in which AL1 and AL2 are the amplitudes (as determined bya nonlinear least squares fit to a double Gaussian function) oflobes 1 and 2 of the DH-PSF, respectively. By measuring (z, LD,LA) from DH-PSF images of an emitter, we are, thus, able todetermine the molecule’s orientation as described below.

SimulationsWe simulated the DH-PSF response to dipole orientation basedon full vectorial diffraction calculations (21), in which dipoleemitters are embedded in a polymer at a fixed distance below theair–polymer interface (SI Text and Fig. S1). The polarized elec-tric field distributions from this calculation were propagated tothe intermediate image plane and convolved with the DH-PSF asin the experimental setup, and the final images (Fig. 2A) were fitwith a double Gaussian estimator (SI Text and Fig. S2). Thisprocedure was repeated for various (z, θ, ϕ) to sample thefunctions LD(z, θ, ϕ) and LA(z, θ, ϕ) at resolution (δz = 50 nm,δθ ∼ 6.5°, δϕ ∼ 6.5°). Fig. 2 B–D shows a z cross-section of this

A

B

C

D

Fig. 1. DH-PSF imaging system. (A) Inverted microscopeand 4f optical system schematic, where L1 and L2 are focallength-matched achromatic lenses. Our sample for theexperiments described consisted of DCDHF-N-6 molecules(Center Inset) embedded in a thin layer of PMMA (Left In-set). Orientation angles (θ, ϕ) are defined in Right Inset andhave ranges (0°, 90°) and (−180°, 180°), respectively. (B) Thehigh efficiency dual-polarization detection DH-PSF setupused for these experiments (inverted microscope omittedfor simplicity). The collected fluorescence is split by a polar-izing beam splitter (PBS) into reflected (R; blue) and trans-mitted (T; red) channels. Input Cartesian unit vectorsðxinput ; y inputÞ define molecular orientation (θ, ϕ) and arepropagated differently through the various reflections inthe two polarization channels [ðxR; yRÞ and ðxT ; yT Þ]. Thetwo electric field polarization axes E

*R and E

*T are projected

identically onto the phase mask (Inset). Inset shows how eachpolarization axis (blue and red arrows) is oriented relative tothe mask’s axis of phase discontinuities (dashed orange)when the mask is upright (i; polarization perpendicular todiscontinuities) and rotated (ii; polarization parallel to dis-continuities). (C and D) Two side-on views of the SLMportion of the setup showing the square pyramidal mirror.

19088 | www.pnas.org/cgi/doi/10.1073/pnas.1216687109 Backlund et al.

Page 3: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

functional behavior, linearly interpolating between samples,whereas Fig. 2 E and F gives two examples of how LA varieswith z for fixed (θ, ϕ). Notably, LA has a distinct functionalform in each of the parallel and perpendicular polarizationchannels (where parallel and perpendicular refer to the relativeorientation between the axis of polarization and the axis of phasediscontinuities in the DH-PSF phase mask) (Fig. 1B Inset) be-cause of the fact that the asymmetry of the phase mask itselfbreaks the degeneracy of the two channels. From the estimatorfit, we also mapped the apparent lateral shifts (Δx, Δy) associatedwith the DH-PSF as a function of (z, θ, ϕ). Fig. 2 G and H showsthat this shift can, indeed, be on the order of ∼200 nm, similar tothe behavior of the standard PSF (10). In general, more highlyinclined molecules (θ closer to 0°) tend to exhibit both larger LAand larger (Δx, Δy).These simulations provided a library that was used to fit ori-

entation and correct the associated position error of a realmeasurement using the following algorithm. First, polarized DH-PSF images were fit with a double Gaussian estimator, yieldingthe observables (xapparent, yapparent, z, LA, LD). The apparentlateral position (xapparent, yapparent) is the true lateral position ofthe molecule (xtrue, ytrue) plus the apparent lateral shift (Δx, Δy)caused by the dipole emission effect. Second, (z, LA, LD) werefed to the simulated look-up table to give an estimate of theorientation (θ, ϕ). Third, (z, θ, ϕ) were referenced by the simu-lation to give a predicted (Δx, Δy) that was then subtracted from(xapparent, yapparent) to recover the true lateral position of themolecule: (xtrue, ytrue). In principle, SM dipole emission alsocauses errors in the double Gaussian-based estimate of z; how-ever, our simulations show that this error is small compared with

our precision in z. Here, we neglect this effect, but in principle, itcan be corrected by extending the scheme above.

Experimental ValidationTo show our ability to fit orientation and correct apparent shifts,we recorded SM images using the setup in Fig. 1B. Samplesconsisted of dicyanomethylenedihydrofuran-N-6 (DCDHF-N-6)(31) molecules spun in a layer of poly(methyl methacrylate)(PMMA) (Fig. 1A Insets) that provided a rigid enough envi-ronment such that angular flexibility of the dipole orientationwas minimized. For a given field of view, the objective wasscanned over z in 50-nm steps over a 2-μm total depth range(DR) centered about the focal plane. Images were recorded atevery z step, and therefore, for each SM, many different sets ofobservables (xapparent, yapparent, z, LA, LD) were recorded. Fromeach single measurement of this set, we estimated orientation and3D position according to the prescription detailed above, and ul-timately, we subtracted lateral shifts (Δx, Δy) from the apparentpositions. Because we do not expect the orientation or lateralposition of an SM to change on our imaging timescale for PMMAat room temperature, each independent measurement shouldproduce the same (xtrue, ytrue, θ, ϕ) within some precision. In otherwords, our method is validated if the determined (xtrue, ytrue, θ, ϕ)of an SM are each constant functions of z.Simulations show that the DH-PSF dipole response is de-

pendent on E⇀-field polarization, as a consequence of the asym-

metry of the phase mask. Because of the geometry of our setup,however, the polarization axis of each polarization channel isidentical in the SLM plane (propagation of E

*R and E

*T in Fig. 1

B–D). This property of our optical system has the effect ofrendering both experimental polarization channels with eitherparallel-type behavior or perpendicular-type behavior, depend-ing on the orientation of the mask (Fig. 1B Inset). Thus, tocapture the full behavior of the DH-PSF response to dipoleemission patterns, we measured each SM with the mask orientedupright (perpendicular) and rotated by 90° (parallel). It is im-portant to note that the two simultaneously recorded images (inthe T and R channels) are not identical and do not purvey de-generate information, despite both exhibiting parallel-/perpen-dicular-type behavior, because the molecular coordinates areprojected differently onto the mask in the T and R channels.Hence, we used nondegenerate information provided from fourdifferent images (two acquisitions of two polarization channels)of each SM to produce a single estimate of (xtrue, ytrue, θ, ϕ). Wecollected ∼3,000–8,500 (Table S1) total photons per set of fourimages, a number on the same order as typical SMACM meas-urements, but the high stability of the SMs allowed this collectionto be done for many z positions.As an independent verification of these orientation estimates,

we also measured orientation directly through defocused imagesof SMs using the standard PSF (21). The defocused images wereacquired by toggling off the DH-PSF phase mask (thereby in-voking a clear aperture) and defocusing the microscope objectiveby 1.00 ± 0.15 μm away from the sample. By comparing theseimages with simulations using template matching (SI Text andFig. S3), we extracted a separate estimate of (θ, ϕ). Interestingly,we found it necessary to correct primary astigmatism and comausing the SLM and include spherical aberration in our simulationsto match experimental images to simulated ones (SI Text). DH-PSF orientation estimation, however, did not require accountingfor these aberrations explicitly to produce the results describedbelow [although some amount is included implicitly in the cali-brated DH-PSF response of (x, y) vs. z] (SI Text). Because the DH-PSF mask itself works by imparting a sizeable distortion on thewave front, the associated images seem to be more robust to minordisturbances of the wave front caused by aberrations (32).

Results and DiscussionUsing our DH-PSF–based method, we estimated the orientationsof six SMs (two example molecules are shown in Fig. 3, and moreexamples are shown in Fig. S4). To distinguish the four images,

z = −0.5 μm z = 0 μm z = 0.5 μm

−1 0 1−1

0

1

sin(θ)cos(φ)

sin(θ )

sin(φ)

−1

0

1

−1

0

1

Lobe

asym

met

ry

−1 0 1−200

0

200

z (μm)

xy s

hift

(nm

)

−1

0

1

Lobe

asym

met

ry

−1 0 1−200

0

200

z (μm)

xy s

hift

(nm

)

A B C

D

FE

HG

Fig. 2. Simulated behavior of the DH-PSF response to dipole orientation.(A) Example DH-PSF images of a molecule with orientation (θ = 45°, ϕ = 180°)at several z positions. Upper (red) shows images that appear in the parallelpolarization channel, whereas Lower (blue) shows images from the per-pendicular polarization channel (definitions in the text). (Scale bar: 1 μm.) (B)A z cross-section (z = −500 nm) of LD as a function of (θ, ϕ), where (θ, ϕ) areprojected into rectangular coordinates according to the relations marked onthe axes. The center of the plot corresponds to a dipole aligned with theoptical axis (θ = 0°), whereas the perimeter of the plot corresponds tomolecules with θ = 90°; ϕ is the azimuthal angle from the positive x axis thatincreases in a counterclockwise direction. The points marked with ○ and ◇correspond to the orientations (θ = 45°, ϕ = 180°) and (θ = 60°, ϕ = −120°),respectively. (C and D) Corresponding plots showing the functional behaviorof LA vs. orientation for constant z = −500 nm in the parallel (C; red axes)and perpendicular (D; blue axes) channels, respectively. (E) LA vs. z in theparallel (red) and perpendicular (blue) channels for a fixed example dipoleorientation (○). (F) The same plot for a different orientation (◇). (G) Δx(solid line) and Δy (dashed line) vs. z in the two channels for the ○ orien-tation. (H) The same plot for the ◇ orientation.

Backlund et al. PNAS | November 20, 2012 | vol. 109 | no. 47 | 19089

APP

LIED

PHYS

ICAL

SCIENCE

S

Page 4: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

we assigned colors to each mask orientation/polarization channelcombination: mask parallel/transmitted polarization is the redchannel, mask perpendicular/transmitted polarization is the goldchannel, mask parallel/reflected polarization is the green channel,and mask perpendicular/reflected polarization is the blue channel.Fig. 3 A and C shows representative images of the DH-PSF foreach example molecule in the four channels at a single z position,whereas Fig. 3 B and D shows the corresponding clear-aperturedefocused images. For both example molecules, we show eachmeasurement of LD (Fig. 3 E and F) and LA (Fig. 3 G and H) asa scatter point plotted vs. z. For each quartet of LA measure-ments and associated LD measurements, there is a correspond-ing estimation of (θ, ϕ). The mean orientation of the Gaussian fitof these distributions of (θ, ϕ) yields the solid overlays in Fig. 3E–H. The full distributions of DH-PSF–extracted (θ, ϕ) areshown in Fig. 3 I and J. The sequential build up of these histo-grams as the objective was scanned for molecule 1 is shown inMovie S1. Also displayed for each molecule in Fig. 3 I and J isthe orientation estimated from our independent defocusedmeasurement (Fig. 3 I and J, purple arrows). We show excellentagreement between the defocus-determined orientation and theDH-PSF–based measurements: (θDH-PSF = 42° ± 12°, ϕDH-PSF =−76° ± 7°) and (θdefocus = 40° ± 2°, ϕdefocus = −81° ± 4°) for mol-ecule 1; (θDH-PSF = 61° ± 2°, ϕDH-PSF = 141° ± 4°) and (θdefocus =63° ± 3°, ϕdefocus = 143° ± 5°) for molecule 2. The SDs of eachdistribution (2°–12°) are comparable with the SDs of othermethods (12).

Apparent shift corrections for these example molecules areshown in various ways in Fig. 4. First, (Δx, Δy) are plotted asfunctions of z and overlaid again with those predictions for themean orientation fit (Fig. 4 A–D). By binning the (x, y) positionsof both SMs recorded across our entire 2-μm DR, we producedthe 2D histograms shown in Fig. 4 E and F. For an isotropicemitter, for which (xapparent, yapparent) does not depend on z, the2D distribution should be circularly symmetric, with width ap-proximately proportional to 1=

ffiffiffiffi

Np

, where N is the number ofphotons collected in that channel. Because of the dipole effect,the uncorrected distributions have irregular, elongated shapes insome cases (e.g., the green channel of molecule 1 and the greenand red channels of molecule 2). By subtracting the apparentshifts (Δx, Δy) predicted from each individual fit of orientation,we recovered the corrected distributions shown in Fig. 4 G andH. Importantly, cases that were elongated and irregularly shapedwhen uncorrected became more concentrated and symmetricwhen corrected. Uncorrected cases that were relatively concen-trated and symmetric to begin with did not undergo much changeupon correction (e.g., the gold channel of molecule 2). The goldchannel of molecule 1 shows a case where a large (>200 nm) shiftis followed closely by simulation and removed on correction, but

phot

ons/

pixe

l

0

50

100

20 40 600

102030

θ (deg)

Num

ber

−100 −500

102030

φ (deg)40 60 80

0100200

θ (deg)

Num

ber

120 140 1600

50100150

φ (deg)

−1

0

1

Line

ardi

chro

ism

−1

0

1

Line

ardi

chro

ism

−1 −0.5 0 0.5 1−1

0

1

z (μm)

Lobe

asym

met

ry

−1 −0.5 0 0.5 1−1

0

1

z (μm)

Lobe

asym

met

ry

A B C D

FE

HG

JI

Fig. 3. Orientation fitting results for example molecule 1 (A, B, E, G, and I)and example molecule 2 (C, D, F, H, and J). (Scale bars: 1 μm.) (A) Four ex-ample DH-PSF images (which constitute one measurement of orientation) ofmolecule 1 at z ∼ −250 nm as it appears in each of four mask orientation/polarization channel combinations: red, gold, green, and blue (definitions inthe text). Note that only one lobe is easily visible in the red and goldchannels. (B) The transmitted (Upper) and reflected (Lower) polarizationdefocused standard PSF images used for independent orientation mea-surement. (C) The equivalent of A for example molecule 2 at z ∼ 0 nm. (D)The equivalent of B for example molecule 2. (E and F) Each measurement ofLD (scatter points) and the predicted LD based on the mean fit orientation(solid line) for each molecule. (G and H) Each measurement of LA in eachchannel (color code is the same as in A) and the overlaid predicted LA for themean fit orientation. (I and J) Histograms of θ and ϕ extracted from DH-PSFbased measurements. Magenta line is Gaussian fit. Purple arrow denotesorientation extracted from defocused imaging.

−200

0

200Δx

(nm

)

−1 −0.5 0 0.5 1−200

0

200

z position (μm)

Δy (n

m)

a

x

y

a a

0

5

10

15

20

0 0.5 1 1.5 210

30

50

Depth range (μm)

σ a (n

m)

−200

0

200

Δx (n

m)

−1 −0.5 0 0.5 1−200

0

200

z position (μm)

Δy (n

m)

a

x

y

a a

0

10

20

30

0 0.5 1 1.5 210

30

50

Depth range (μm)

σ a (n

m)

BA

DC

FE

HG

JI

LK

Fig. 4. Measured and predicted (Δx, Δy) caused by dipole orientationeffects for example molecule 1 (A, C, E, G, I, and K) and molecule 2 (B, D, F, H,J, and L). Color code is the same as in Fig. 3. (A–D) Measured (scatter points)Δx and Δy vs. z for each example molecule, with overlay (solid line) ofsimulated shift based on the mean extracted orientation. For each molecule,one mask orientation/polarization combination did not produce meaningfullocalizations because of low signal and/or high LA (omitted channels). (E andF) 2D histogram of uncorrected (xapparent, yapparent) localizations over the2-μm DR. In each panel, a is the predominant direction of lateral shift forthat mask orientation/polarization channel. Bin size, 15 nm. (G and H) Cor-responding 2D histograms of the corrected localizations as produced bysubtracting each predicted (Δx, Δy) based on each individual estimation of(θ, ϕ). (I and J) The corrected 2D histograms produced by subtracting thepredicted (Δx, Δy) based on the average estimation of (θ, ϕ). Displayed (x, y)axes are 100 nm in length. (K and L) Additional quantification of the im-provement in lateral localization showing σa, the SD along the direction a ineach channel, as a function of DR about the focal plane. We compareσa calculated for the uncorrected case (solid line), the individual measurement-based correction (dashed line), and the average-based correction (dotted line).

19090 | www.pnas.org/cgi/doi/10.1073/pnas.1216687109 Backlund et al.

Page 5: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

fewer photons detected in this channel make for a more diffusecorrected histogram than in other cases.Large values of LA can also affect the precision of position

measurements, because the midpoint between one very brightand one very dim lobe is difficult to localize. Commonly in ourexperiment, a combination of few detected photons and largeLA made spatial localizations in one of four mask orientation/polarization channels highly erratic (omitted cases in each ex-ample molecule). However, we still obtained meaningful esti-mates of LD and LA from such cases that were highly valuable inthe extraction of (θ, ϕ). These orientation estimates, in turn,resulted in shift corrections that improved the localizations inchannels that did have meaningful (x, y, z) fits. There is, thus,a tradeoff between the benefit of being able to fit orientation andcorrect dipole-induced shifts and the cost of splitting photonsinto multiple channels.For additional improvement in correction, Fig. 4 I and J shows

the results of using the average values of all measurements of(θ, ϕ) to correct each localization. In general, this method im-proved the localizations slightly more than the individual mea-surement-based method, because it makes use of a less noisycorrection vector. The individual measurement-based method ismore practical, however, because it requires far fewer measure-ments and therefore, is easier to adapt for use in a SMACMmeasurement. Taken together, these observations suggest thatan individual set of four images is sufficient to correct the bulk ofthe dipole-induced shift for a molecule but that this correctioncan be improved somewhat if it is possible to measure the samemolecule several times at multiple z positions.As a final quantification, we also calculated the SD of (x, y)

localizations along the lateral direction a for each channel andmolecule over various ranges in z (Fig. 4 K and L). The axis a foreach case is depicted in the lower left of each panel in Fig. 3 Eand F, and it was determined by treating the 2D distribution asan ellipse and finding its major axis. Thus, a large σa implies asystematic apparent shift along the direction a. A correcteddistribution should remove the systematic shift and therefore,should reduce σa to the SD expected from photon-limited pre-cision. Data for a given DR about the focal plane were calculatedby including all localizations for which jzj was less than that DR/2. As expected, in each case for which the deviation along a wasrelatively large, the corrected values were markedly improved,especially at large DR. Table S2 summaries the σa values calcu-lated at the full 2-μm DR and compares them with the σ expectedfrom photon-limited precision. For example, in the green chan-nel of molecule 1, the uncorrected σa (54 nm) was three timeslarger than the statistical localization precision (18 nm). Theindividual measurement-based correction improved σa to 35 nm,whereas the average-based method improved it further to 24 nm.In the gold channel of molecule 1, the large (>200 nm) shiftproduced an uncorrected σa of 116 nm (vs. 28-nm precisionalong that direction). Individual correction reduced this numberby more than a factor of two (55 nm), whereas the averagecorrection brought σa to within 6 nm of the precision (34 nm)along that direction. This channel is still relatively diffuse alongthe x direction on correction, however, because of limited photondetection (the DH-PSF in general gives unequal σx and σy, be-cause it is not circularly symmetric) (27). In the green channel ofmolecule 2, σa (48 nm) was nearly two times as large as theprecision (25 nm). This SD was corrected to within 5 nm of theprecision in both the individual correction (30 nm) and the av-erage correction (28 nm). Similarly, the red channel of molecule2 gave uncorrected, individually corrected, and average correctedσa values of 35, 21, and 18 nm compared with a precision of 17 nm.In some cases (e.g., the red channel of molecule 2), we found

that the corrected value was actually slightly worse at small DRbut still much better at sufficiently high DR. This effect suggeststhat an efficient algorithm that corrects orientation effects ina SMACM experiment will only apply corrections in cases inwhich it is beneficial (e.g., at large jzj, molecules with inclinedorientations, and polarization channels with a high enough

signal-to-background ratio). Overall, these results show that theDH-PSF has the powerful ability to correct large lateral positionerrors caused by the SM dipole effect over an extended 2-μm DR.

Conclusion and OutlookWith this direct experimental demonstration, we show that theDH-PSF can be used to simultaneously extract precise 3D lo-calization, estimate dipole orientation, and dramatically reduce(x, y) systematic errors caused by the orientation effect over anextended z range. Although our method requires two cameraexposures for every set of measurements, our SLM can alternatephase mask orientations programmatically at speeds of at least30 Hz (limited by the liquid crystal composition of our SLM).Different phase modulators can be toggled faster. Additionally,an optical setup containing separate phase masks for each po-larization channel may be able to provide orientation from justa single acquisition using a slightly modified analysis.This proof of principle shows that our method works best for

correcting shifts of intermediately inclined molecules (∼θ ∈ [35°,75°]). Our method has more difficulty fitting the orientations ofless-inclined molecules (θ > 75°); because jLAj is closer to zerofor all z, there exist near-degeneracies in some of these cases(Fig. S4). However, this limitation is not likely to prohibit thecorrection of significant dipole-induced mislocalizations in SMACMexperiments, because those same noninclined molecules producenegligible (Δx, Δy) (Fig. S5). We did not encounter many veryhighly inclined molecules (θ < 35°) in our measurement, becauseboth pumping and collection efficiencies are diminished for thesecases (33). A standard SMACM experiment also would havedifficulty detecting these molecules for the same reasons.The theoretical limit of the DH-PSF’s ability to extract posi-

tion and orientation can be quantified using a Fisher informationcalculation; in particular, because the double-lobed shape of theDH-PSF is conserved over various dipole orientations and axialpositions, the orientation precision of the DH-PSF is uniformover a large range of (z, θ, ϕ) (SI Text and Fig. S6). Thus, becausethe DH-PSF has been established as a highly precise method for3D SMACM (28, 29), our orientation extraction/shift correctionmethod to improve accuracy by removing systematic error isa good candidate to be incorporated into such experiments. Aspointed out above, if labels are rotationally mobile, the dipoleshift can be averaged out during an acquisition. At the otherextreme, if labels are fixed in orientation during an acquisition,our method can be applied. The intermediate regime of rota-tional flexibility will be the subject of future work. Importantly,the fact that DH-PSF LA deviates from zero when SMs are fixedin their orientation and sufficiently inclined suggests that theDH-PSF can be used as a diagnostic tool in determining whenlabels are sufficiently rotationally mobile.Although we found that we did not need to explicitly account for

aberrations in the DH-PSF to yield good results for the examplesgiven here, it may be necessary to address aberrations in the futureto broaden the scope of our method and improve its performance.Finally, the DH-PSF–based orientation-sensing method may alsobe improved by using more sophisticated estimators and phaseretrieval (34), particularly when high LA or aberrations make theDH-PSF deviate from its typical double Gaussian shape. With thecorrection of localization errors from the SM dipole orientationeffect, far-field superresolution microscopy is one step closer toattaining molecular spatial resolution (∼1 nm) (35) to reveal thenanoscale machinery at work within living cells.

Materials and MethodsSample Preparation. Nanomolar concentrations of DCDHF-N-6 (31) were spunin a thin layer of 1% (by mass) PMMA of thickness 30–35 nm as measured byellipsometry. Fluorescent beads (FluoSpheres, 100-nm diameter, 580/605; Invi-trogen) were adhered directly to the air–polymer interface by spinninga dilute solution on top of the polymer layer and allowing it to dry. Thebeads served as both internal calibration markers for the DH-PSF responseand fiducial markers for drift and sample tilt corrections (SI Text). They could

Backlund et al. PNAS | November 20, 2012 | vol. 109 | no. 47 | 19091

APP

LIED

PHYS

ICAL

SCIENCE

S

Page 6: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

not be spun directly into the polymer matrix because of instability inorganic solvents.

Imaging. Samples were mounted on an inverted Olympus IX71 fluorescencemicroscope, and the optical setup described in Fig. 1B was constructed outsidethe left-side port with the use of a phase-only spatial light modulator (BoulderNonlinear Systems XY Phase Series). Fluorophores were excited with an Ar-ionlaser emitting at 514 nm. The wide-field excitation beam was tilted at a slightangle to compensate for the dynamic range of brightness presented by thebeads vs. the SMs. The laser was roughly circularly polarized (∼1.4:1) at thesample. Molecules were irradiated at relatively low intensity (∼0.1 kW/cm2) toensure a full cycle of measurements before photobleaching. Fluorescence wascollected through a 100× 1.4 N.A. oil-immersion objective (Olympus UPlan-SApo 100×/1.40) and filtered using a Chroma Z514RDC dichroic and a 590/60band pass. Images were recorded on an Andor iXon+ DU897-E EMCCD cameraoperating at an EM gain setting of 300.

For DH-PSF imaging, the sample was scanned over a z range of 2 μm at 50-nm spacing using an objective z positioner (C-Focus operating in open-loopmode; Mad City Labs). The DH-PSF mask was first loaded on the SLM withperpendicular orientation. Frames were recorded with 0.5-s exposures at5 frames per z step. The loaded mask was then rotated 90° clockwise, and thesample was scanned again. After two full scans, the sample was refocusedand then defocused by 1.00 ± 0.15 μm; therefore, the objective was movedaway from the air–polymer interface. The coma/astigmatism correction mask(SI Text) for the transmitted channel was then loaded onto the SLM, and 10–30 1-s exposures were recorded. Then, the coma/astigmatism correction maskfor the reflected channel was loaded, and another such series of images wasrecorded. This full process of scanning two times and then taking defocusedsnapshots was repeated until a field of view was sufficiently photobleached.

Analysis. Tiff image stacks were exported by the Andor Solis software andanalyzed with custom MATLAB routines. Each analyzed molecule was pickedby hand based on the presence of a single-step bleaching event (indicating anSM), sufficient local sparsity to avoid overlapping signals, and sufficientsignal-to-noise ratio. A constant background offset was subtracted fromeach SM region, which was calculated from a user-defined local backgroundregion. The resulting image was fit to a double Gaussian (i.e., the sum of two

Gaussians) through nonlinear least squares regression using the MATLABfunction lsqnonlin (SI Text). Simultaneously recorded gold and blue channelimages (mask perpendicular) of an SM were paired with green and redchannel images (mask parallel) of the same molecule based on z proximity.From each set of four images, the observables (xapparent, yapparent, z, LA, LD)were calculated. (xapparent, yapparent) were given by the midpoint positionbetween the two lobes as determined by the double Gaussian fit along witha composite calibration correction to account for the nonideality of the DH-PSF response because of optical aberrations, sample tilt, stage drift, and thedifference between the surrounding medium of the SMs vs. the medium ofthe fluorescent beads (SI Text and Fig. S7); z was given by the measuredangle of the line connecting the centers of the two lobes relative to thehorizontal, which was referenced to the calibrated response of the bead DH-PSF to the stepping of the stage by known amounts. LA was computed fromthe estimated amplitudes of the two Gaussians. LD was calculated from thetotal integrated photons above background in the region of the SM. If theautomatic double Gaussian fit failed, the image was fit to two Gaussians byhand-selecting the regions of the two peaks. In this way, bounds or esti-mates of LA and LD could still be extracted and fed to the orientation look-up program even if (x, y, z) localizations were not reliable in that channel. Toestimate (θ, ϕ) for each measurement of (z, LD, LA), the MATLAB functionlsqnonlin was used to converge on the best (θ, ϕ) that minimized the dif-ference between the measured LA and LD vs. predictions of the theoreticalresponse of the DH-PSF (SI Text).

ACKNOWLEDGMENTS. We acknowledge Dr. Michael Thompson andDr. Majid Badieirostami for assistance designing the experiments and writingsimulation code. We also thank Dr. Sean Quirin for helpful discussions andMaxShulaker for ellipsometry measurements. M.P.B. acknowledges supportfrom a Robert and Marvel Kirby Stanford Graduate Fellowship. M.D.L.acknowledges support from a National Science Foundation Graduate Re-search Fellowship and a 3Com Corporation Stanford Graduate Fellowship.A.S.B. acknowledges support from the National Defense Science and Engi-neering Graduate Fellowship. This work was supported by National ScienceFoundation Awards DBI-0852885 (to R.P.), DBI-1063407 (to R.P.), and DGE-0801680 (to R.P.), and by National Institute of General Medical SciencesGrant R01GM085437 (to W.E.M.).

1. Moerner WE (2012) Microscopy beyond the diffraction limit using actively controlledsingle molecules. J Microsc 246(3):213–220.

2. Hell SW (2009) Microscopy and its focal switch. Nat Methods 6(1):24–32.3. Betzig E, et al. (2006) Imaging intracellular fluorescent proteins at nanometer reso-

lution. Science 313(5793):1642–1645.4. Hess ST, Girirajan TPK, Mason MD (2006) Ultra-high resolution imaging by fluores-

cence photoactivation localization microscopy. Biophys J 91(11):4258–4272.5. Rust MJ, Bates M, Zhuang X (2006) Sub-diffraction-limit imaging by stochastic optical

reconstruction microscopy (STORM). Nat Methods 3(10):793–795.6. Sharonov A, Hochstrasser RM (2006) Wide-field subdiffraction imaging by accumu-

lated binding of diffusing probes. Proc Natl Acad Sci USA 103(50):18911–18916.7. Ha T, Enderle T, Chemla S, Selvin R, Weiss S (1996) Single molecule dynamics studied

by polarization modulation. Phys Rev Lett 77(19):3979–3982.8. Dickson RM, Norris DJ, Moerner WE (1998) Simultaneous imaging of individual molecules

aligned both parallel and perpendicular to the optic axis. Phys Rev Lett 81(24):5322–5325.9. Enderlein J, Toprak E, Selvin PR (2006) Polarization effect on position accuracy of

fluorophore localization. Opt Express 14(18):8111–8120.10. Engelhardt J, et al. (2011) Molecular orientation affects localization accuracy in su-

perresolution far-field fluorescence microscopy. Nano Lett 11(1):209–213.11. Gould TJ, et al. (2008) Nanoscale imaging of molecular positions and anisotropies. Nat

Methods 5(12):1027–1030.12. Rosenberg SA, Quinlan ME, Forkey JN, Goldman YE (2005) Rotational motions of

macro-molecules by single-molecule fluorescence microscopy. Acc Chem Res 38(7):583–593.

13. Sosa H, Peterman EJG, Moerner WE, Goldstein LSB (2001) ADP-induced rocking of thekinesin motor domain revealed by single-molecule fluorescence polarization mi-croscopy. Nat Struct Biol 8(6):540–544.

14. Peterman EJG, Sosa H, Goldstein LSB, Moerner WE (2001) Polarized fluorescencemicroscopy of individual and many kinesin motors bound to axonemal microtubules.Biophys J 81(5):2851–2863.

15. Huang B, Wang W, Bates M, Zhuang X (2008) Three-dimensional super-resolutionimaging by stochastic optical reconstruction microscopy. Science 319(5864):810–813.

16. Juette MF, et al. (2008) Three-dimensional sub-100 nm resolution fluorescence mi-croscopy of thick samples. Nat Methods 5(6):527–529.

17. Shtengel G, et al. (2009) Interferometric fluorescent super-resolution microscopy re-solves 3D cellular ultrastructure. Proc Natl Acad Sci USA 106(9):3125–3130.

18. Pavani SRP, et al. (2009) Three-dimensional, single-molecule fluorescence imagingbeyond the diffraction limit by using a double-helix point spread function. Proc NatlAcad Sci USA 106(9):2995–2999.

19. Ha T, Laurence TA, Chemla DS, Weiss S (1999) Polarization spectroscopy of singlefluorescent molecules. J Phys Chem B 103(33):6839–6850.

20. Peterman EJ, Sosa H, Moerner WE (2004) Single-molecule fluorescence spectroscopyand microscopy of biomolecular motors. Annu Rev Phys Chem 55:79–96.

21. Patra D, Gregor I, Enderlein J (2004) Image analysis of defocused single-moleculeimages for three-dimensional molecule orientation studies. J Phys Chem A 108(33):6836–6841.

22. Lieb MA, Zavislan JM, Novotny L (2004) Single-molecule orientations determined bydirect emission pattern imaging. J Opt Soc Am B 21(6):1210–1215.

23. Sick B, Hecht B, Novotny L (2000) Orientational imaging of single molecules by an-nular illumination. Phys Rev Lett 85(21):4482–4485.

24. Toprak E, et al. (2006) Defocused orientation and position imaging (DOPI) of myosinV. Proc Natl Acad Sci USA 103(17):6495–6499.

25. Mortensen KI, Churchman LS, Spudich JA, Flyvbjerg H (2010) Optimized localizationanalysis for single-molecule tracking and super-resolution microscopy. Nat Methods 7(5):377–381.

26. Aguet F, Geissbühler S, Märki I, Lasser T, Unser M (2009) Super-resolution orientationestimation and localization of fluorescent dipoles using 3-D steerable filters. OptExpress 17(8):6829–6848.

27. Thompson MA, Lew MD, Badieirostami M, Moerner WE (2010) Localizing and track-ing single nanoscale emitters in three dimensions with high spatiotemporal resolu-tion using a double-helix point spread function. Nano Lett 10(1):211–218.

28. Lew MD, et al. (2011) Three-dimensional superresolution colocalization of intracellularprotein superstructures and the cell surface in live Caulobacter crescentus. Proc NatlAcad Sci USA 108(46):E1102–E1110.

29. Lee HL, Sahl SJ, Lew MD, Moerner WE (2012) The double-helix microscope super-resolves extended biological structures by localizing single blinking molecules in threedimensions with nanoscale precision. Appl Phys Lett 100(15):153701.

30. Pavani SRP, DeLuca JG, Piestun R (2009) Polarization sensitive, three-dimensional,single-molecule imaging of cells with a double-helix system. Opt Express 17(22):19644–19655.

31. Lord SJ, et al. (2007) Photophysical properties of acene DCDHF fluorophores: Long-wavelength single-molecule emitters designed for cellular imaging. J Phys Chem A111(37):8934–8941.

32. Ghosh S, Quirin SA, Grover G, Piestun R, Preza C (2012) Double helix PSF engineeringfor computational fluorescence microscopy imaging. Proc SPIE 8227:82270F.

33. Sick B, Hecht B, Wild UP, Novotny L (2001) Probing confined fields with single mol-ecules and vice versa. J Microsc 202(Pt 2):365–373.

34. Quirin S, Pavani SRP, Piestun R (2012) Optimal 3D single-molecule localization forsuperresolution microscopy with aberrations and engineered point spread functions.Proc Natl Acad Sci USA 109(3):675–679.

35. Pertsinidis A, Zhang Y, Chu S (2010) Subnanometre single-molecule localization,registration and distance measurements. Nature 466(7306):647–651.

19092 | www.pnas.org/cgi/doi/10.1073/pnas.1216687109 Backlund et al.

Page 7: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

Supporting InformationBacklund et al. 10.1073/pnas.1216687109SI TextImage Formation of an Electric Dipole Emitter. Our simulationsmodel the simplified optical system depicted in Fig. S1, which issimilar to systems depicted in refs. 1–3. Specifically, we consideran emitter embedded slightly below the surface of a polymer thatis spin-coated on a glass coverslip. In this situation, the objectivelens will collect forward propagating light emitted by the dipolein addition to backward propagating light that has been reflectedby the air–polymer interface. Because the refractive index mis-match between the polymer and glass is minor (nglass = 1.518 vs.npolymer = 1.49), we ignore the effects of this interface. After theelectric fields at the image plane have been calculated, it isstraightforward to augment our simulations to account for polar-ization, aberrations, and the Double-Helix Point Spread Function(DH-PSF) mask.Theoretical calculations that extend the pioneering work of

Richards and Wolf (4) and predict the image formed by anemitter with a fixed dipole orientation embedded at or near aninterface have been developed in the studies by Enderlein (5)and Böhmer and Enderlein (6). In addition, ref. 7 describesmethods for simulating a freely rotating emitter or a cluster oftightly packed randomly oriented dipoles (such as a nanoscalefluorescent bead). Furthermore, refs. 8 and 9 describe the impact ofsubwavelength thin films on dipole emission. Subsequent inves-tigations (10) showed the effects of layered media on the in-tensity distributions observed at the back focal plane andimage plane of a high-N.A. optical system.Our general strategy will be to decompose the emission pattern

of the dipole into a basis of plane waves. The effects of the air–polymer interface and the image-forming optics will be determinedfor each plane–wave component, and the intensity distributionformed on a camera sensor will then be calculated by integratingover all of the plane–wave contributions that propagatethrough the imaging system. We begin by considering a dipoleemitter embedded a distance z0 below the surface of an air–polymer interface, with a fixed azimuthal orientation ϕ = α anda polar angle of θ = β. From ellipsometry measurements, wedetermined the polymer thickness z0 to be 30–35 nm. If weconsider the plane–wave component of this dipole’s emission withpropagation direction given by k ¼ fsinðηÞcosðψÞ; sinðηÞsinðψÞ; cosðηÞg, where ψ and η are the azimuthal and polar ori-entations of the propagation direction, respectively, then theelectric field associated with this wave is given by

E�

η;ψ� ¼ es

h

sin�

β�

sin�

ψ − α�

Eks

η�

i

þ eph

sin�

β�

cos�

ψ − α�

Ekp

η�þ cos

β�

E⊥p

η�

i

; [S1]

where

Eks�

η� ¼ − nglass

e−ikz0 cosðηÞ þ Rs�

η�

eikz0 cosðηÞ�

Ekp�

η� ¼ nglass cos

�

e−ikz0 cosðηÞ −Rp�

η�

eikz0 cosðηÞ�

E⊥p

η�¼ nglass sin

�

e−ikz0 cosðηÞ þ Rp�

η�

eikz0 cosðηÞ�

: [S2]

In the above expressions, es ¼ f− sinðψÞ; cosðψÞ; 0g and ep ¼fcosðψÞcosðηÞ; sinðψÞcosðηÞ; − sinðηÞg are the unit vectors per-pendicular to k, with es also perpendicular to the optical axis.The superscripts k and ⊥ denote whether a given component ofthe E-field is parallel or perpendicular to the interface. Thewave number is k ¼ 2πnglass=λ, and Rs(η) and Rp(η) are the

Fresnel reflection coefficients for s and p polarized waves at apolymer–air interface (these terms account for the portion ofbackward-propagating light reflected at the interface); λ = 609nm for DCDHF-N-6 in poly(methyl methacrylate) (11). Theeffect of the optical imaging system is to map a plane wave withpropagation direction k to another plane wave with a new propa-gation direction k′ ¼ fsinðη′Þcosðψ ′Þ; sinðη′Þsinðψ ′Þ; cosðη′Þg. Thismapping is determined by Abbe’s sine condition: M sin(η′) =nglass sin(η); here, we assume an image formed in air with mag-nification M. The azimuthal coordinate is unchanged betweeninput and output (i.e., ψ ′= ψ). Finally, to determine the electricfield present at the image plane, all angular contributions to thetotal field are integrated together. Closed-form expressions existfor the integration over ψ ′; however, the integration over η′ mustbe done numerically. The following integral must be evaluated:

Exðρ;ϕÞEy�

ρ;ϕ�

¼Z n′max

0

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

cosðη′Þnglasscos

η�

s

e−ikz1 cosðηÞ sin�

η′�

exey

dη′: [S3]

Note that the image plane has been parameterized using polarcoordinates (ρ, φ) and that the polarized components of theelectric field are calculated for two orthogonal directions, xand y. In the above expression, the upper bound of integrationis set by the maximally inclined plane wave that can be capturedby the microscope objective, and it is determined by the objective’snumerical aperture (NA = 1.4) and magnification (M = 100):η′max ¼ sin−1

NA=M� ¼ :014 rad. Furthermore, the amount of

defocus is given by z1. [Focusing beyond the emitter (i.e., movingthe objective to the coverslip) corresponds to z1 of positive sign.]The terms ex and ey are given by

ex ¼ sinðβÞ2

n

J0 cos�

α�

h

cos�

η′�

Ekp −Ek

s

i

− J2 cos�

2φ− α�

h

cos�

η′�

Ekp þ Ek

s

io

− iJ1 cos�

β�

cos�

φ�

cos�

η′�

E⊥p

ey ¼ sinðβÞ2

n

J0 sin�

α�

h

cos�

η′�

Ekp −Ek

s

i

− J2 sin�

2φ− α�

h

cos�

η′�

Ekp þ Ek

s

io

− iJ1 cos�

β�

sin�

φ�

cos�

η′�

E⊥p ;

[S4]

where J0,1,2 are Bessel functions of the first kind with the argu-ment k′ρ sin(η′), and k′¼ 2π

λ . Similar expressions exist for themagnetic fields. However, we assume that, for all of the propa-gation media, μr = 1. For nonpolarization sensitive detection, theintensity distribution at the image plane may be calculated as

I�

ρ;φ� ¼ cε0

2

h

Ex�

ρ;φ�

Ex*�

ρ;φ�þ Ey

ρ;φ�

Ey*�

ρ;φ�

i

: [S5]

Practically, the integral in Eq. S3 was computed using a rectan-gular approximation by substituting Eq. S2 into Eq. S4 and thenevaluating the integrand at a fixed location (ρ, φ) while varying η′from zero to η′max. Summing the results together yields the elec-tric fields at the single point in the image plane (ρ, φ). In our

Backlund et al. www.pnas.org/cgi/content/short/1216687109 1 of 10

Page 8: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

simulations, Ex(ρ, φ) and Ey(ρ, φ) were evaluated within a 256 ×256 grid, with samples spaced a distance of 1 μm apart. (Afteraccounting for the magnification factor of 100, the effective spac-ing at the focal plane is 10 nm.) Also, for each location withinthe image plane, the integrand was evaluated for 100 differentvalues of η′ spaced evenly between zero and η′max (i.e., δη′ ¼:01× sin−1

1:4=100� ¼ :00014 rad). Performing this calculation

for an entire 256 × 256 region takes approximately 2 min run-ning 32-bit MATLAB on a standard office desktop computer(Dell Optiplex 960 with a 3 GHz Intel Core 2 Duo CPU).Equipped with a theoretical model describing the image

formed by a dipole emitter, we may begin to make modificationsto account for the particulars of our optical design. Because apolarizing beam splitter was used to record two orthogonalpolarizations of the electric field in separate regions of an imagesensor, the two polarized intensity distributions [Ix(ρ, φ) andIy(ρ, φ)] are calculated as

Ixðρ;φÞIy�

ρ;φ�

¼ cε02

"

Ex

ρ;φ

Ex*

ρ;φ

Ey

ρ;φ

Ey*

ρ;φ

#

: [S6]

Furthermore, the geometry of our setup causes light reflected bythe beam splitter to incur one additional reflection as it propa-gates through the imaging pathway. Therefore, simulated imagesfor one polarization must be reflected about the φ = 90° axisbefore comparison with acquired data. (Here, the x and y polarizedsimulations can be related to the T and R channels, respectively,by taking into account the appropriate reflections and rotationsintroduced by the experimental setup. Fig. 1B depicts the specificgeometrical transformations that each polarized image under-goes before it is projected onto the image sensor.)

Modeling Dipole Emission Effects on the DH-PSF. To simulate how adipole emitter undergoes phase modulation to form the DH-PSF,we use the same framework described above and numericallyintegrate Eq. S3. The electric fields were sampled with a differentpixel size and number of samples from above to match the discretepixels of the DH-PSF phase mask and avoid ringing artifactsof the discrete Fourier transform (1,024 × 1,024 pixels, 5.95-μmpixel size in the image plane, 59.5-nm pixel size in the focal plane).The resulting electric fields at the image plane are calculated as

EDH−PSFx

ρ;φ� ¼ FT

n

FTfExðρ;φÞgeiψDH −PSFðρ′;φ′Þo

EDH−PSFy

ρ;φ� ¼ FT

n

FTfEyðρ;φÞgeiψDH −PSFðρ′;φ′Þo : [S7]

In the above equation, FT{} denotes the Fourier transform op-eration, whereas two cascaded Fourier transforms model theimage formation process of a 4f optical system (12). The functionψDH-PSF(ρ′,φ′) models the phase delay imparted by the DH-PSFpattern programmed into the spatial light modulator (SLM). Inthis simulation, the mask is oriented such that its discontinuitiesare along the x axis; therefore, the x polarization has parallel-type behavior, whereas the y polarization has perpendicular-typebehavior. It is convenient to normalize the radial coordinate, ρ′,such that the electric field is zero at distances greater than ρ′ = 1from the center of the pupil. (If one were to use nonnormalizedcoordinates, ρ′ = 1 corresponds to the distance f NA=M from thecenter of the aperture, where f is the focal length of the optical sys-tem.) Simulated polarized images of the DH-PSF (Fig. S2 A and B),expressed as IDH−PSF

x ∝ jEDH −PSFx j2 and IDH−PSF

y ∝ jEDH −PSFy j2,

were then fit by our double Gaussian estimator to map out theresponse of DH-PSF as a function of dipole orientation (θ, ϕ).The double Gaussian estimator finds the best-fitting parame-

ters ðAL1;L2; xL1;L2; yL1;L2; σL1;L2Þ that minimize the summedsquare error between the function

IDGðρ;φÞ ¼ AL1 exph− ðρ cosφ− xL1Þ2 − ðρ sinφ− yL1Þ2

2σ 2L1

i

þ AL2 exph− ðρ cosφ− xL2Þ2 − ðρ sinφ− yL2Þ2

2σ 2L2

i

[S8]

and the simulated DH-PSF images IxDH-PSF and Iy

DH-PSF usingthe optimization function lsqnonlin. ðAL1;L2; xL1;L2; yL1;L2; σL1;L2Þare the fitted amplitudes, x position, y position, and widths (SD),respectively, of each Gaussian lobe of the DH-PSF. The doubleGaussian estimator produced reasonable fits of the DH-PSF formost dipole orientations (Fig. S2 C and D). The spatio-orienta-tion space (z, θ, ϕ) was sampled at a rate of (δz = 50 nm, δθ ≅6.5°, δϕ ≅ 6.5°), which is shown as black dots in Fig. S2E, column1. Observable parameters of the DH-PSF for orientation fittingwere then calculated from the double Gaussian fit parametersusing the relations

LDðz; θ;ϕÞ ¼ Nx −Ny

Nx þ Ny; [S9]

LAðz; θ;ϕÞ ¼ AL1 −AL2

AL1 þ AL2; [S10]

Δxðz; θ;ϕÞ ¼ xL1 þ xL22

; [S11]

and

Δyðz; θ;ϕÞ ¼ yL1 þ yL22

; [S12]

where Nx and Ny are the total number of photons in the x and ypolarization channels, respectively. Linear dichroism LD wasfound to have essentially no dependence on emitter defocus z(Fig. 2B), but lobe asymmetry LA (Fig. S2 E and F) and lateralshifts (Δx, Δy) (Fig. S2 G–J) have diverse behavior across (z, θ,ϕ). Notably, it is the contrasting behavior between LA in the xand y channels [termed the parallel (Fig. S2E) and perpendicular(Fig. S2F) behaviors in the text] that breaks the degeneracy ofLD and enables effective measurement of orientation with theDH-PSF.

Modeling and Correction of Optical Aberrations. Our objective lensand relay optics induced a slight amount of spherical aberration inthe images recorded on the electron-multiplying charge-coupleddevice (EMCCD) camera. Simulations of the standard PSF wereaugmented accordingly by applying a spherical aberration phasemask to the electric field that we calculated to be present at thepupil (aperture) plane of our microscope. We found the aberratedelectric fields for x and y polarizations simply by taking theFourier transform of the unaberrated fields present at the imageplane, multiplying by the appropriate phase mask, and then in-verse Fourier transforming:

ESph:Ab:x ðρ;φÞ ¼ IFT

n

FTfExðρ;φÞgeiψSph:Ab;ðρ′;φ′Þo

ESph:Ab:y ðρ;φÞ ¼ IFT

n

FT�

Ey�

ρ;φ��

eiψSph:Ab:ðρ′;φ′Þo : [S13]

IFT{} denotes the inverse Fourier transform operation. The ab-erration function ψSph.Ab.(ρ′, φ′) models the phase delay effects of

Backlund et al. www.pnas.org/cgi/content/short/1216687109 2 of 10

Page 9: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

spherical aberration. Using the ρ′ coordinate system, first-orderspherical aberration is given explicitly by the polynomial (13)

ψSph:Ab:�ρ′;ϕ′� ¼ A

6ρ′4 − 6ρ′2 þ 1�

: [S14]

The coefficient A sets the magnitude of aberration present in thesystem. We determined this coefficient heuristically. Interestingly,the incorporation of spherical aberration into the simulated im-ages only improved matches with the clear-aperture defocusedPSFs, and therefore, it was only included there. Accounting forspherical aberration did not improve matches to simulated DH-PSF images and therefore, was omitted for this part of the analysis.The clear-aperture defocused PSFs also exhibited elements of

astigmatism and comatic aberration. Instead of incorporatingthese effects in simulation, we corrected these distortions ex-perimentally using the same SLM that was used to create the DH-PSF. Because coma and astigmatism are inherently asymmetricand lead to more dramatic distortions in the defocused PSF, wefound it more expedient to experimentally diagnose and removethese aberrations. In-house software was developed for loadingcoma and astigmatism phase masks onto the SLM and adjustingthe magnitude of their aberration coefficients. We used phasemasks of the form

ψComa�

ρ′;φ′� ¼ B

3ρ′3 − 2ρ′�

cos�

φ′−φComa�

ψ Astig:�

ρ′;φ′� ¼ C

6ρ′2cos�

2�

φ′−φAstig:��� : [S15]

To find optimal values for the parameters B, C, φComa, and φAstig.,data from our EMCCD was displayed in real time, and the SLMdriving software was used to modify the aberration coefficients tominimize irregularities in the PSF. Because of the unique geom-etry of our optical setup, the SLM appears to have differentorientations when viewed in each polarization channel. Further-more, the imaging pathways for the two different polarizationsinduced slightly different aberrations. Therefore, it was neces-sary to image the x and y polarized images of defocused mole-cules sequentially, with different phase masks loaded onto theSLM to correct the aberrations present in each of the separateimaging paths. When recording DH-PSF data, we observed thatthe 3D PSFs were more robust to the effects of minor aberra-tions, and no experimental corrections were required.

Single-Molecule Orientation and 3D Position Estimation with the DH-PSF. As discussed in Materials and Methods, (xapparent, yapparent, z,LA, LD) was measured for each single molecule (SM) in eachdetection channel (red, gold, green, and blue in the text) byfitting SM DH-PSF images with a double Gaussian estimator.This estimator is identical to the one used for the simulatedDH-PSF response and is detailed above. SM orientation andposition measurements were completed in two steps. First, ori-entation was estimated using the measured z position, lobe asym-metry, and linear dichroism across the four detection channels.Next, given this orientation, the calculated (Δx, Δy) shift fromsimulations was subtracted from (xapparent, yapparent) to yield thetrue lateral position of the SMs.The orientation measurement process was carried out as fol-

lows. Calibrated measurements of (z, LA, LD) from each channelwere grouped together such that their measured z positions werewithin 50 nm of each other. (Channels with a double Gaussian fitthat failed because of severe lobe asymmetry had their lobeasymmetry measured by hand-designating each Gaussian spotof the DH-PSF and fitting it to a 2D Gaussian function). Themeasurements (zmeas,i, LAmeas,i, LDmeas,i) from all channels iwere input to the orientation-fitting algorithm, which usedthe MATLAB function lsqnonlin to find the orientation (θ, ϕ)that minimizes

X

4

i¼1

n

wðLAsimðzi; θ;ϕÞÞ�

LAmeas;i −LAsimðzi; θ;ϕÞ�2

þ �LDmeas;i −LDsimðzi; θ;ϕÞ�2o

;

[S16]

the weighted squared difference between measurements and sim-ulations of LA and LD, where the index i refers to measurementsfrom each of the four channels,

wðxÞ ¼8

>

>

<

>

>

:

1; jxj< 2 =

3

3ð1− jxjÞ; jxj≥ 2 =

3

; [S17]

the subscripts meas and sim refer to experimental and theoreticalvalues, respectively, and zi = zmeas,i + zoffset,i is a calibrated zposition that compensates for relative defocus between polariza-tion channels of the 4f system. The piecewise linear-weightingfunction w(x) serves to deemphasize simulated values of LA thatare difficult to measure because of low signal-to-noise ratio (i.e.,LA is deemphasized when jLAj> 2 =

3). Explicitly, LDmeas,i is cal-culated from pairs of channels such that

LDred ¼ LDgreen ¼ Nred −Ngreen

Nred þ Ngreen[S18]

and

LDgold ¼ LDblue ¼ Ngold −Nblue

Ngold þ Nblue: [S19]

Finally, zoffset,i is calibrated for each measured SM by finding thebest value that allows measurements of LA to overlap with cal-culated values of LA for the entire z scan. Because of localsample tilt, this value is slightly different for each SM that wemeasured. For molecule 1, zoffset = 70 nm for the transmittedpolarization channels, and zoffset = 200 nm for the reflected po-larization channels. For molecule 2, zoffset = 0 nm for the trans-mitted polarization channels, and zoffset = −200 nm for thereflected polarization channels. For molecule 3, zoffset,T = 186nm, and zoffset,R = −185 nm. For molecules 4, 5, and 6, zoffset,T =zoffset,R = 0 nm. For positions and orientations (zi, θ, ϕ) wheresimulations were not explicitly carried out, the MATLAB in-terpolation function TriScatteredInterp was used to evaluateLAsim(zi, θ, ϕ) and LDsim(zi, θ, ϕ). The above orientation esti-mator is a first-order approach that matches experimental datato a library of simulated data to measure orientation; more so-phisticated optimization functions or finer sampling of simulateddata may improve orientation-fitting performance.When matching experimental measurements of (z, LA, LD)

to simulations, care must be taken in ensuring that ϕ istransformed appropriately considering the optical system andorientation of the DH-PSF mask. Mapping ϕIIP in the in-termediate image plane (IIP) (Fig. 1B) to a specific orienta-tion relative to the DH-PSF mask (ϕmask), we obtain thefollowing relations:

ϕred;mask ¼ ϕIIPϕgold;mask ¼ ϕIIP þ π=2ϕgreen;mask ¼ 3π=2−ϕIIPϕblue;mask ¼ −ϕIIP

: [S20]

Furthermore, the red and green channels show parallel-type be-havior (Fig. 1B Inset) because their polarization is parallel to the

Backlund et al. www.pnas.org/cgi/content/short/1216687109 3 of 10

Page 10: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

phase discontinuities of the DH-PSF mask, whereas the gold andblue channels show perpendicular-type behavior.Two strategies for calculating the true lateral position of each

SM were used. The first strategy uses measurements of (zmeas,i,LAmeas,i, LDmeas,i) at a single z position from all four channels i tomeasure orientation (θ, ϕ). Then, the calculated [Δxi(zi, θ, ϕ),Δyi(zi, θ, ϕ)] shift from simulations is subtracted from the ap-parent position (xapparent,i, yapparent,i) for each channel to recoverthe true location of the molecule. The second strategy involvesaveraging all such pointwise measurements of orientation overthe entire z scan to yield ðθ;ϕÞ. Then, the calculated shift[Δxiðzi; θ;ϕ;Þ;Δyiðzi; θ;ϕÞ] for this single orientation ðθ;ϕÞ is usedto correct the lateral position of the SM. This approach has thebenefit of reducing orientation measurement noise at the cost ofrequiring multiple snapshots of each SM. Again, the MATLABinterpolation function TriScatteredInterp was used to evaluate[Δxi(zi, θ, ϕ), Δyi(zi, θ, ϕ)].

Clear-Aperture Defocused Image Template Matching. By matchingsimulations to the image of a defocused molecule (withoutthe DH-PSF), a well-established SM orientation determinationmethod,wegain an independent estimate of thatmolecule’s dipoleorientation. This information is used for verification of our DH-PSF–based orientation estimation algorithms. Data were com-pared with a list of simulated templates. The dipole orientationof a defocused molecule was then estimated as the orientation ofthe template that yielded the closest fit to the actual data.Using the simulation methods described above, a list of

templates was generated for both the x and y polarized imagesof a defocused molecule. This template list included dipoleemitters simulated at all orientations of θ from 0° to 90° and ϕfrom 0° to 355° spaced at intervals of 5°. Hence, 1,368 pairs ofpolarized intensity distributions were simulated. In practice,all dipole emission patterns were calculated using MATLAB.Spherical aberration was simulated by first calculating theunaberrated high-resolution electric fields of an emitter withfixed orientation at the image plane. The result was zero-padded to ensure adequate sampling in the Fourier domain,and then, it was discrete Fourier-transformed using MAT-LAB’s fft2 function. The appropriate phase masks were ap-plied as described above, and then, an inverse Fouriertransform was applied using MATLAB’s ifft2 function to re-cover the aberrated image. By visually comparing simulationswith acquired data, we found that the spherical aberrationcoefficient A = 1.8 yielded simulated images that closelymatched experiment (Eq. S14). Furthermore, by comparingsimulation with experiment, we estimated the defocus to be1 ± 0.15 μm below the emitter. (For a given field of view, theprecise defocus depth was estimated by eye to generate a setof templates). To account for the pixelation effects of theEMCCD, simulations were performed on a high-resolutiondiscretization of the image plane (256 × 256-μm extent with1 × 1-μm sampling, which is equivalent to 2.56 × 2.56 μm with10 × 10-nm pixels in object space). Subdivisions were made inthe high-resolution grid (16 × 16 μm corresponding to 160 ×160 nm in object space), and intensity values within eachsubdivision were summed together. This summation leads to16 × 16-pixel templates that match the actual pixel size of ourEMCCD. Each pair of polarized images was then normalizedby the sum of the squares of its pixel values [that is, if Ijx andIjy denote the x and y polarized images corresponding withthe jth template in our list, the normalized templates arecalculated as

~Ijx ¼

Ijxffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

P

M

m¼1

P

M

n¼1

Ijx½m; n��2þ�Ijy½m; n��2s

~Ijy ¼

Ijyffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

P

M

m¼1

P

M

n¼1

Ijx½m; n��2þ�Ijy½m; n��2s

: [S21]

The additional brackets are used to denote the [m,n]th pixel in animage, and M is the total number of pixels along a given dimen-sion in the template (in this case, M = 16).The normalized template that most closely matched the

defocused images was determined as follows. A 16 × 16 region ofinterest was drawn around the image made by both the x and ypolarized intensities (T and R channels, respectively, of Fig. 1B)of an SM. Background intensity was estimated by calculating themean number of photons per pixel detected in a manually se-lected rectangle of ∼20 × 20 to 40 × 40 pixels nearby the mol-ecule of interest and subsequently subtracted from each pixel inthe region of interest containing the single defocused molecule.The cross-correlation between each of the two background-subtracted polarized images, Dx and Dy, and each template ofgenerated images was then computed. The maximum values ofthe two resulting cross-correlation matrices were summed to-gether, and the result, Cj, was stored. Mathematically, we eval-uated

Cj ¼ maxp;q

X

M

m¼1

X

M

n¼1

Dx�

m; n�

~Ijx

mþ p; nþ q�

!

þmaxp;q

X

M

m¼1

X

M

n¼1

Dy�

m; n�

~Ijy

mþ p; nþ q�

!

: [S22]

This operation can be performed efficiently using MATLAB’sxcorr2 function. After Cj was computed for the entire templatelist, the template yielding the largest value of Cj was chosen asthe best match, and the dipole orientation of that template wasused as an estimate of the dipole orientation of the data. Fig. S3shows the results of using template matching to estimate dipoleorientation for the two molecules examined in the text. Repre-sentative image acquisitions are plotted with the simulated im-ages of the templates that yield the best match. For bothmolecules, a defocus of z1 = 1.13 μm was used to generate a listof templates. To benchmark the precision of this technique, ori-entation was estimated repeatedly for the same molecule usingdata from successive acquisitions. The mean orientation estimatesand SDs in θ and ϕ were calculated (Results and Discussion).

Cramer–Rao Lower Bound for the Polarization-Sensitive DH System.In the text,wepresented thepolarization-sensitiveDHmicroscopeas ameans to simultaneously measure 3D position and orientationof an SM. Here, we quantitatively compare this approach withestablishedmethods that use a clear-aperture standardPSF system(1, 7). We calculate the photon-limited precision that can beachieved in estimating the position and orientation of a dipoleusing the Cramer–Rao Lower Bound (CRLB) (14).The CRLB is defined as the inverse of the Fisher Information

(FI) matrix, which assesses the information contained in a prob-ability distribution for estimation of parameters. The FI matrix isadditive; therefore, for a multiple channel system, the FI matrix ofeach channel is summed to get the total FI and then inverted toget the CRLB. The lower bound SD (σLB), which is the squareroot of the CRLB, directly yields a lower bound for the precisionof an unbiased estimator in the same units as the measured data.

Backlund et al. www.pnas.org/cgi/content/short/1216687109 4 of 10

Page 11: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

Thus, for the five-parameter estimation problem, the σ matrix isa 5 × 5 matrix. For 3D imaging and localization, we are in-terested in the minimum localization volume. One measure ofthis uncertainty volume is

σ3D ¼ 4π3σx·σy·σz: [S23]

For the CRLB simulation, the dipole is assumed to be im-mersed in a medium of index 1.52 (no interface) and emitting atλ = 610 nm. The objective lens has an N.A. of 1.4. The pixel sizeis 160 nm in the object plane. We assume the imaging system tobe shift-invariant, which is a good approximation in the centralregion of the field of view. The number of photons collected bythe objective lens depends on the dipole’s polar orientation.CRLB calculations show that most photons are collected whenthe dipole is oriented along θ = 90°, and thus, for photon-limitedsystems, all images are normalized according to this case (samenumber of photons emitted). The total number of photons de-tected for the dipole along θ = 90° is taken to be 5,000, and allcalculations are performed in the shot noise limit. In all calcu-lations, the number of photons detected in the standard system isthe same as the number in the DH system properly dividedamong the four channels.Fig. S6A shows the localization and orientation error lower

bound as a function of θ with and without background. Theamount of defocus for either system is chosen to optimize theprecision based on the lowest CRLB. Thus, the axial defocus forthe standard (clear aperture) system is z = 100 nm, whereas theDH system performs best at focus (z = 0). Our calculations alsoshow that σ3D, σθ, and σϕ are relatively constant as a function ofthe polar angle θ. To make a fair comparison, we choose ϕ = 45°in Fig. S6A to evenly distribute the light between the two linearpolarization channels.The lower bound of the error for angle estimation and 3D

position using the four-channel polarization-sensitive DH systemis substantially lower than the lower bound of the standard PSFcase. As expected, with the inclusion of background, the locali-zation precision of the DH worsens, but it still performs betterthan the standard case for intermediate θ. The standard systemperformance is calculated for the optimal defocus, and therefore,it will substantially deteriorate with variations in z. On the con-trary, the DH system has a slower variation with defocus andhence, provides a relatively uniform performance with defocus.The lower bound for estimating the angle and position asa function of defocus is shown in Fig. S6 B and C. The dipole isoriented along a representative direction (θ, ϕ) = (45°, 45°) inFig. S6B and (θ, ϕ) = (90°, 90°) in Fig. S6C. It can be seen thatthe DH system performs better than the standard system overmost of the defocus range with and without background.

Calibrations. The behavior of the DH-PSF vs. z was calibratedusing a fiducial fluorescent bead. Each field of view was chosensuch that it included at least one fiducial bead near the edge of

the field. Laser power was kept sufficiently low such that thebead did not bleach appreciably during the experiment, whichwould otherwise make its emission increasingly anisotropic. LDof the beads was found to be ∼0 ± 0.1, justifying its use as anisotropic emitter. The microscope objective was then scanned inz as described in Materials and Methods. The angle made by thetwo lobes of the DH-PSF image of the bead at each z step wasrecorded to produce an angle vs. z calibration curve to be appliedto the fit SMs (Fig. S7A), where z = 0 was assigned to 0°. Al-though the simulated DH-PSF of an isotropic point emitter doesnot translate in (x, y) as a function of z, factors such as sampletilt, aberrations, and imperfect mask alignment produce a base-line shift that must also be calibrated out of SM localizations torecover the true dipole-induced shifts (Fig. S7 B and C). Becausesome of the factors that contribute to these calibration curvesare different in each of the four mask orientation/polarizationchannel combinations, we used unique calibration curves for eachchannel (red, gold, green, and blue in the text).Sample drift is another experimental effect that could poten-

tially swamp the dipole-induced shifts that we sought to measure.By measuring a separate bead-based calibration curve for eachindividual z scan of each field of view, we mitigated this potentialsource of error. In this way, any (x, y) drift occurring duringa scan was included in the (x, y) vs. z calibration. We measuredmultiple reproducible (x, y) shifts across all of our SM z scans,suggesting that drift was effectively removed.The nanoscale behavior of the DH-PSF is subtly affected by the

medium surrounding the SMs and fluorescent beads. Asexplained in Materials and Methods, sample constraints pre-cluded the ability to embed the beads in the same medium as theSMs. Consequently, we measured nonzero residual (x, y) shiftsfor SMs with θ near 90° (we expect these shifts to be constant),even after applying the bead-based calibrations explained above.To calibrate and remove this residual shift, we chose two mol-ecules with predicted shifts that were fairly constant measuredexperimentally to have polar orientations θ of ∼80° and ∼85°.Calibration molecule 1 gave good signal in the reflected polari-zation channel, and therefore, we used the deviation between itsmeasured (x, y) shifts and its predicted shifts to correct all SMshifts in the perpendicular/reflected (blue) and parallel/reflected(green) channels. Calibration molecule 2 gave good signal in thetransmitted polarization channel, and therefore, it was usedequivalently for the perpendicular/transmitted (gold) and par-allel/transmitted (red) channels. We then applied third–eighth-order polynomial fits (chosen by using the lowest order that re-produced the major features as judged by eye) to these residualcurves, effectively smoothing them. In cases where the curvesneeded to be extrapolated to span a full 2-μm depth of field, thesmoothed curves were extended with a linear or low-orderpolynomial that gave the most qualitatively reasonable extrapo-lation. The measured residual shifts and their piecewise poly-nomial fits are displayed in Fig. S7 D and E. These curves weresubtracted from the measured shifts of all molecules in the ap-propriate channels.

1. Patra D, Gregor I, Enderlein J (2004) Image analysis of defocused single-molecule imagesfor three-dimensional molecule orientation studies. J Phys Chem A 108(33):6836–6841.

2. Toprak E, et al. (2006) Defocused orientation and position imaging (DOPI) of myosinV. Proc Natl Acad Sci USA 103(17):6495–6499.

3. Enderlein J, Toprak E, Selvin PR (2006) Polarization effect on position accuracy offluorophore localization. Opt Express 14(18):8111–8120.

4. Richards B, Wolf E (1959) Electromagnetic diffraction in optical systems. II. structure of theimage field in an aplanatic system. Proc R Soc Lond A Math Phys Sci 253(1274):358–379.

5. Enderlein J (2000) Theoretical study of detection of a dipole emitter through anobjective with high numerical aperture. Opt Lett 25(9):634–636.

6. Böhmer M, Enderlein J (2003) Orientation imaging of single molecules by wide-fieldepifluorescence microscopy. J Opt Soc Am B 20(3):554–559.

7. Mortensen KI, Churchman LS, Spudich JA, Flyvbjerg H (2010) Optimized localizationanalysis for single-molecule tracking and super-resolution microscopy. Nat Methods7(5):377–381.

8. Lukosz W (1981) Light emission by multipole sources in thin layers. I. radiationpatterns of electric and magnetic dipoles. J Opt Soc Am 71(6):744–754.

9. Hellen EH, Axelrod D (1987) Fluorescence emission at dielectric and metal-filminterfaces. J Opt Soc Am B 4(3):337–350.

10. Axelrod D (2012) Fluorescence excitation and imaging of single molecules neardielectric-coated and bare surfaces: A theoretical study. J Microsc 247(2):147–160.

11. Lord SJ, et al. (2007) Photophysical properties of acene DCDHF fluorophores: Long-wavelength single-molecule emitters designed for cellular imaging. J Phys Chem A111(37):8934–8941.

12. Goodman JW (2005) Introduction to Fourier Optics (Roberts & Company, GreenwoodVillage, CO).

13. Mahajan VN (2007) Optical Shop Testing, ed Malacara D (Wiley, New York), pp498–546.

14. Kay SM (1993) Fundamentals of Statistical Signal Processing: Estimation Theory(Prentice Hall, Englewood Cliffs, NJ).

Backlund et al. www.pnas.org/cgi/content/short/1216687109 5 of 10

Page 12: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

ψ

φ

η

ρ

k

focal plane

single molecule

air-polymer interface

z0z

optical system

image plane

yx

z

η'

k'

sample

ψ'

I(ρ,φ)

1

(ψ, η)(ψ , η )''

θΦ

m

Fig. S1. The simplified imaging system modeled by our simulations. An SM is laterally located along the optical axis, embedded a distance z0 beneath thesurface of a polymer, and positioned z1 above the focal plane with respect to the optical system. The dipole orientation of the molecule is defined by the polarand azimuthal angles θ and ϕ, respectively. The light that propagates through the optical system can be expressed as a sum of plane waves with wave vectorkðψ ; ηÞ, where ψ and η are the azimuthal and polar orientations of the propagation direction, respectively. Our simulations determine how a given input kðψ ; ηÞis mapped to a different plane wave k’ðψ ’; η’Þ at the output of the optical system. The resulting intensity distribution I(ρ, φ) at a given point in the image plane(parameterized using polar coordinates) is then determined by integrating the contributions of each output plane wave component.

z = −0.6 μm

DH

−PS

F im

age

z = −0.6 μm

DG

fit i

mag

e

−1

0

1

Lobe

asy

mm

etry

sin(θ)

sin(φ)

sin(θ)cos(φ)

−1

0

1

Δx (n

m)

sin(θ)

sin(φ)

sin(θ)cos(φ)

−1 0 1−1

0

1

Δy (n

m)

sin(θ)

sin(φ)

sin(θ)cos(φ)−1 0 1

sin(θ)cos(φ)

z = −0.2 μm z = −0.2 μm

sin(θ)cos(φ)

sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

z = 0.2 μm z = 0.2 μm

sin(θ)cos(φ)

sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

z = 0.6 μm z = 0.6 μm

sin(θ)cos(φ)

sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

−1 0 1sin(θ)cos(φ)

−1

0

1

−200

0

200

BA

DC

FE

HG

JI

−200

0

200

Fig. S2. DH-PSF response to SM orientation and axial position. Simulated DH-PSF images of an SM with orientation (θ, ϕ) = (47°, −173°) at several z positions(each column) in the (A) x (parallel) polarization channel (red) and (B) y (perpendicular) polarization channel (blue). (C and D) Images of double Gaussian fits ofthe DH-PSF images in A and B, respectively. (E and F) DH-PSF lobe asymmetry behavior as measured by the double Gaussian estimator as a function of ori-entation in the two polarization channels. Orientation space is plotted as a 2D map, with horizontal coordinates given by sin(θ)cos(ϕ) and vertical coordinatesgiven by sin(θ)sin(ϕ). Black dots in E show the various SM orientations that were simulated to build the library of DH-PSF behaviors; (G and H) x shift of theDH-PSF in the two polarization channels resulting from the dipole emission pattern, and (I and J) y shift of the DH-PSF in the two polarization channels. In E–J,◇ denotes the orientation (θ, ϕ) = (47°, −173°). (Scale bars: 1 μm.)

Backlund et al. www.pnas.org/cgi/content/short/1216687109 6 of 10

Page 13: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

pho

tons

/pix

el

10

30

50

10

30

50

70

pho

tons

/pix

el

A B

C D

E F

G H

Fig. S3. Determining orientation from defocused images. A and B show representative acquisitions of molecule 1 in the T and R channels, respectively. C andD show the most closely matching set of polarized templates. E and F show T and R channel images of molecule 2, whereas G and H depict the best templates.For these two molecules, a defocus of z1 = 1.13 μm was used to generate a set of templates. For molecule 1, the mean orientation fit and SD were based on 72separate 1-s acquisitions. For molecule 2, 27 acquisitions were collected before photobleaching. (Scale bars: 1 μm.)

40 60 800

10

20

30

θ (deg)

Num

ber

60 80 1000

10

20

30

φ (deg)

Num

ber

−1 0 1−1

0

1

z (μm)

Line

ar d

ichr

oism

Molecule 3

−1 0 1−1

0

1

z (μm)

Lobe

asy

mm

etry

z(θ

DH−PSF=52°, φ

DH−PSF=85°)

(θdefocus

=55°, φdefocus

=84°)

40 60 800

20

40

60

θ (deg)

−80 −60 −400

20

40

φ (deg)

−1 0 1−1

0

1

z (μm)

Molecule 4

−1 0 1−1

0

1

z (μm)

z(θ

DH−PSF=69°, φ

DH−PSF=−59°)

(θdefocus

=70°, φdefocus

=−60°)

40 60 800

20

40

60

θ (deg)

120 140 1600

20

40

60

φ (deg)

−1 0 1−1

0

1

z (μm)

Molecule 5

−1 0 1−1

0

1

z (μm)

z(θ

DH−PSF=70°, φ

DH−PSF=143°)

(θdefocus

=85°, φdefocus

=145°)

60 80 1000

20

40

θ (deg)

100 120 1400

20

40

φ (deg)

−1 0 1−1

0

1

z (μm)

Molecule 6

−1 0 1−1

0

1

z (μm)

z(θ

DH−PSF=75°, φ

DH−PSF=121°)

(θdefocus

=85°, φdefocus

=55°)

Molecule 1

z(θ

DH−PSF=42°, φ

DH−PSF=−76°)

DH−PSF=52°, φ

DH−PSF=85°)

(θdefocus

=55°, φdefocus

=84°)

z(θ

DH−PSF=69°, φ

DH−PSF=−59°)

(θdefocus

=70°, φdefocus

=−60°)

z(θ

DH−PSF=70°, φ

DH−PSF=143°)

(θdefocus

=85°, φdefocus

=145°)

z(θ

DH−PSF=75°, φ

DH−PSF=121

(θdefocus

=85°, φdefocus

=55°)

Molecule 1 Molecule 2xy

z(θ

DH−PSF=61°, φ

DH−PSF=141°)

y x xy xy xy

xy

φ

θ

Fig. S4. Orientation estimations of additional molecules. In the top two rows, each measurement of LD and LA (scatter points) along with the predicted LDand LA (solid curve) based on the mean fit orientation of each molecule is shown. In the middle two rows, histograms of DH-PSF–based orientation estimationsare overlaid with a Gaussian fit (magenta curve). Defocused clear-aperture estimates of orientation are denoted by purple arrows. In the bottom two rows,molecular orientation diagrams depict the measured orientations of molecules 3–6 as well as the orientations of molecules 1 and 2 from the text for com-parison. Both molecules 3 and 4 have intermediate inclination, θ ∈ (35°, 75°), and therefore, they are well fit by the DH-PSF method. Molecules 3 and 4 showexcellent agreement between the orientations estimated by the DH-PSF and the orientations estimated from defocused image template matching. Molecules 5and 6 are not inclined much at all (θ > 75°), and therefore, the DH-PSF method does not yield orientation estimates that agree as well. Small values of jLAj leadto degeneracies in the estimated parameters for these molecules. However, for the purpose of correcting dipole-induced mislocalizations, this fact is not ofmuch importance, because noninclined molecules have negligible shifts (Fig. S5).

Backlund et al. www.pnas.org/cgi/content/short/1216687109 7 of 10

Page 14: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

Δx (n

m)

Molecule 3

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

Δy (n

m)

2D lo

caliz

atio

n hi

stog

ram

s

0

5

10

15

20

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

Molecule 4

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

0

5

10

15

20

25

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

Molecule 5

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

0

5

10

15

20

25

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

Molecule 6

−1 −0.5 0 0.5 1−200

−100

0

100

200

z position (μm)

0

5

10

15

20

25

Fig. S5. Measured and predicted (Δx,Δy) caused by dipole effect for molecules 3–6. Measured (scatter points) Δx (Top) and Δy (Middle) vs. z for each molecule(each column), with overlay (solid line) of predicted shift based on the mean extracted orientation. Highly inclined molecules (molecules 3 and 4; columns 1 and2) generally experience larger dipole shifts than slightly inclined molecules (molecules 5 and 6; columns 3 and 4). For molecules 3–5, one mask orientation/polarization combination did not produce meaningful localizations because of low signal and/or high LA (omitted channels). Images in Bottom show 2Dhistograms of uncorrected (xapparent,yapparent) localizations (Top), corrected localizations based on individual estimations of (θ, ϕ) (Middle), and corrected lo-calizations based on average estimate of (θ, ϕ) (Bottom). Histograms for cases with significant directional shifts (green channel of molecules 3 and 4; bluechannel of molecule 4) become more concentrated and less elongated when corrected. Bin size, 15 nm. (Scale bars: 100 nm.) Color code is the same as in Figs. 3and 4.

0 500

20

40

θo

σ φ (

o )

Bckgrnd=5

0 50

1

2

3

θo

σ θ (

o )

0 50100

200

300

400

θo

σ 3D (

nm3 )

0 500

20

40

θo

σ φ (

o )

Bckgrnd=0

0 500.5

1

1.5

2

θo

σ θ (

o )

0 50

50

100

150

200

θo

σ 3D (

nm3 )

−1 0 1

1.5

2

2.5

3

z (μm)

σ φ (

o )

Bckgrnd=5

−1 0 10

5

10

z (μm)

σ θ (

o )

−1 0 1

500

1000

1500

z (μm)

σ 3D (

nm3 )

−1 0 1

0.5

1

1.5

z (μm)

σ φ (

o )

Bckgrnd=0

−1 0 10

2

4

z (μm)

σ θ (

o )

−1 0 1

100200300400500

z (μm)

σ 3D (

nm3 )

−1 0 1

1

2

z (μm)

σ φ (

o )

Bckgrnd=0

−1 0 10.5

1

z (μm)

σ θ (

o )

−1 0 10

500

1000

1500

z (μm)

σ 3D (

nm3 )

−1 0 1

1

2

3

4

z (μm)

σ φ (

o )

Bckgrnd=5

−1 0 1

1

1.5

2

z (μm)

σ θ (

o )

−1 0 10

5000

z (μm)

σ 3D (

nm3 )

A B C

Fig. S6. Lower bound of the precision (SD) for estimating the azimuthal angle (σϕ), polar angle (σθ), and 3D position (σ3D) based on CRLB calculations. Bluesolid curves represent the standard PSF case, and dashed red curves represent the DH case. (A) Precisions as functions of θ without background (Left) and witha background of five photons per pixel (Right). ϕ = 45°, the standard PSF case is defocused by 100 nm, and the DH case is in focus. (B and C) Precisions asfunctions of defocus (z) without background (Left) and with a background of five photons per pixel (Right) for a dipole with (B) orientation (θ, ϕ) = (45°, 45°)and (C) orientation (θ, ϕ) = (90°, 90°).

Backlund et al. www.pnas.org/cgi/content/short/1216687109 8 of 10

Page 15: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

−1 −0.5 0 0.5 1−80

0

80

(noitisopz μm)

angl

e (d

egre

es)

−1 −0.5 0 0.5 1−200

0

200

(noitisopz μm)

resi

dual

Δx

(nm

)

−1 −0.5 0 0.5 1−200

0

200

(noitisopz μm)

resi

dual

Δy

(nm

)

−1 −0.5 0 0.5 1−200

0

200

(noitisopz μm)

base

line

x sh

ift(n

m)

−1 −0.5 0 0.5 1−200

0

200

z position (μm)

base

line

y sh

ift(n

m)

A

B C

D E

Fig. S7. Calibration curves used in SM fitting. Color code is same as in Figs. 3 and 4. (A) Angle made by two DH-PSF lobes vs. z as measured from a fluorescentbead. Different optical paths and different lateral offsets of the phase mask make for different curves in the four channels. (B and C) Baseline x and y vs. zcurves. Nonidealities such as sample tilt, aberrations, and lateral mask offset cause a nonzero lateral shift in the PSF, even for an isotropic emitter, which mustbe subtracted out. (D and E) Residual lateral shifts vs. z. Because the fiducial beads are embedded in a slightly different medium, application of just the beadcalibration curves leaves some residual shift for SMs that is caused by factors other than dipole orientation. Residual lateral shift calibration curves are cal-culated by subtracting the predicted shifts from the apparent shifts of two calibration molecules (SI Text).

Table S1. Average photons detected above background in eachchannel for each molecule

Molecule no. Green Red Gold Blue Total

1 1,100 700 510 1,300 3,6102 1,900 2,600 2,100 2,200 8,8003 2,200 240 820 2,600 5,8604 2,000 1,400 1,200 1,900 6,5005 1,300 2,100 1,800 1,300 6,5006 1,600 930 650 1,100 4,280Mean total 5,925

Calculated by integrating all photons within a 15 × 15-pixel box contain-ing the molecule and subtracting the average background photons containedin this area. Average background for each molecule was determined by mea-suring the background in a nearby hand-designated region (∼9 photons/pixelper frame average across all molecules).

Backlund et al. www.pnas.org/cgi/content/short/1216687109 9 of 10

Page 16: Simultaneous, accurate measurement of the 3D position and ......Simultaneous, accurate measurement of the 3D position and orientation of single molecules Mikael P. Backlunda,1, Matthew

Movie S1. DH-PSF–based molecular orientation measurements of molecule 1 over two axial (z) scans. The raw DH microscope images from each of the fourmeasurement channels (red, gold, green, and blue) are shown in the upper left. Linear dichroism (middle left) and lobe asymmetry (bottom left) are depictedas scatter points because they are measured during the scans. Finally, histograms of the calculated polar and azimuthal orientation of molecule 1 are plotted atright. The final frames of the movie show LD and LA (as solid lines) for the mean orientation measured for molecule 1 (center of the Gaussian fits at right).

Movie S1

Table S2. σa at 2-μm depth range for molecules 1 and 2 compared to localization precision

Green Red Gold Blue

σa (nm) σa/precision σa (nm) σa/precision σa (nm) σa/precision σa (nm) σa/precision

Molecule 1Uncorrected 54 3 — — 116 4.1 36 1.4Individually corrected 35 1.9 — — 55 2 37 1.5Mean corrected 24 1.3 — — 34 1.2 36 1.4Precision 18 — — — 28 — 25 —

Molecule 2Uncorrected 48 1.9 35 2.1 26 0.96 — —

Individually corrected 30 1.2 21 1.2 28 1 — —

Mean corrected 28 1.1 18 1.1 27 1 — —

Precision 25 — 17 — 27 — — —

Precision was calculated by binning localizations into 100-nm z bins and taking the SD along a in each bin. These values were thenaveraged across all bins within the center ∼1 μm of the z range. Our corrections show significant improvement in cases that haveσa ∼2× larger than precision or worse (molecule 1, green and gold channels; molecule 2, green and red channels).

Backlund et al. www.pnas.org/cgi/content/short/1216687109 10 of 10