比特派钱包app官方下载网址|dsf

作者: 比特派钱包app官方下载网址
2024-03-07 18:21:30

无损音乐格式APE/FLAC/WAV/DFF/DSF有什么不同?_音频

无损音乐格式APE/FLAC/WAV/DFF/DSF有什么不同?_音频

新闻

体育

汽车

房产

旅游

教育

时尚

科技

财经

娱乐

更多

母婴

健康

历史

军事

美食

文化

星座

专题

游戏

搞笑

动漫

宠物

无障碍

关怀版

无损音乐格式APE/FLAC/WAV/DFF/DSF有什么不同?

2020-07-31 13:39

来源:

QCJPSZS

原标题:无损音乐格式APE/FLAC/WAV/DFF/DSF有什么不同?

1、 前三种都是属于俗称的“无损”音频文件格式。

APE和FLAC的文件相比WAV体积要小一些,在音频的提取和转化过程中,也经过一定的压缩,因此和原始录音对比,APE和FLAC文件本质也是“有损”,只不过损失很小,接近可以忽略不计;其中APE算法有较好的压缩率,相比之下FLAC压缩率有所不如,因此同品质的FLAC文件往往略大。

WAV是由微软公司开发的一种电脑音频格式,属于数字化的无损音频格式,诞生较早,也比较常见,一般而言WAV音频的文件体积都较大。

2、 DFF和DSF,都是采用DSD编码的音频格式。

DSD是Direct Stream Digital的缩写,表示直接比特流数字编码,是SACD(Super Audio CD)的编码模式,由Sony与Philips在1996年共同开发诞生,与采用PCM编码的DVD-AUDIO阵营竞争。

DSD编码用1bit比特流的方式取样,采样率2.8224MHz (CD 44.1kHz取样的64倍)的高取样方式,直接把模拟音乐信号波形以脉冲方式转变为数字信号,并以约四倍于传统CD音频文件的空间储存音乐,因此可以提供更为优秀的声音效果。由于取样次数高,所以取样过的波形很圆顺,十分接近模拟波形。

展开全文

DFF和DSF就是采用DSD编码所制作的音频文件,由于DSD由索尼和飞利浦共同开发,因此两家也分别采用了DFF和DSF作为文件格式名,DFF文件最初被索尼用于PS游戏机上的音频,后来进一步的衍生为专用于音乐产品的SACD,并深受发烧友推崇。

DFF和DSF主要用于没经过压缩的双声道录音,体积比WAV文件更大。由于DSD编码的音频文件体积庞大,对于数据量更大的多声道音频,一般会另外采用经过压缩的DST格式。

在网络年代,数字音乐早已成为主流,甚至在汽车音响领域,数字音乐取代传统的CD播放也已成为趋势,而在这个热潮之中,上述无损格式音频,凭借着更佳的音质,也受到了发烧级听众的追捧。然而,要聆听无损音频,无论哪种格式,首先都需要播放设备支持对相关格式的解码,而目前的汽车音响业界,能够全面支持APE,FLAC,WAV,DFF,DSF等格式的高品质音源还是较为稀缺。

近期,英国Gypsy Sound(吉普赛之声)即将推出的首款数字播放器,就会全面支持上述各种无损音频格式,并且在音频的播放品质上,也达到让人惊艳的高水准,对于每一位的Car HI-FI玩家而言,这款数字播放器的面世,相信都将会是2020年里令人难忘的一份巨大惊喜。返回搜狐,查看更多

责任编辑:

平台声明:该文观点仅代表作者本人,搜狐号系信息发布平台,搜狐仅提供信息存储空间服务。

阅读 ()

推荐阅读

一文读懂差示扫描荧光法(DSF) - 知乎

一文读懂差示扫描荧光法(DSF) - 知乎切换模式写文章登录/注册一文读懂差示扫描荧光法(DSF)e测试服务平台e测试,科研资源共享及专业检测与分析平台,专业,高效,让科研更简单。差示扫描荧光法(Differential Scanning Fluorimetry, 简称DSF),是在荧光定量PCR仪上缓慢加热样品,在加热的过程中检测荧光染料与结构发生改变的蛋白质相结合的量,来评价蛋白质热稳定性的方法。该方法具有蛋白样品损耗少、通量高、温度变化范围广及数据准确等优点,被广泛用于蛋白质稳定性(蛋白质热稳定性参数及其影响因素)、蛋白结构和构象、蛋白-配体相互作用及蛋白质稳定剂、抑制剂、辅助因子等领域的研究。01 DSF原理DSF可以通过荧光染料或蛋白内源荧光信号监测升温过程中蛋白构象的变化计算其熔解温度Tm(折叠蛋白与去折叠蛋白相等时的温度)。染料法DSF最常用的是SYPRO Orange染料,SYPRO Orange 是一种环境敏感的疏水染料,当温度升高时,蛋白去折叠,疏水部分暴露出来,染料与蛋白质的疏水部分特异性结合,荧光增强。在有特定化合物或配体结合的情况下,蛋白稳定性会上升,表现为熔解温度的上升。图1. 染料法DSF无标记DSF (Nano DSF)则是基于蛋白去折叠过程中色氨酸发射光谱的位移进行检测。因此,通过检测荧光变化,可实现在非标记环境下评估蛋白质热稳定性、化学稳定性、胶体稳定性、等温稳定性、变复性能力等性质。图2. Nano DSF02 DSF测试过程1、仪器准备DSF 主要是利用荧光定量 PCR 仪在 96 孔或 384 孔 PCR 板上完成平行检测,以实现高通量筛选目的蛋白和化合物。2、样品准备蛋白样品需要较高的纯度(~75%),化合物的纯度通过 LC-MS 质检;SYPRO Orange 是目前DSF实验中应用最广泛的染料,方便且高效。此外,也有一些特殊的染料用于特定蛋白的研究,例如 1,8-ANS、2,6-ANS、2,6-TNS、oxazole 系列和 BODIPY、BFC 系列染料探针。3、反应体系配置根据实验设计把蛋白和配体配置成不同浓度的体系,置于 96 孔板或 384 孔板。缓冲液可以根据实际需要进行更换。DSF 的反应体系主要包括蛋白样品、反应缓冲液、荧光染料及候选的化合物配体,每检测孔样品体积为 20~25 µL,以蛋白缓冲液作为空白对照,每个检测组设三组重复。4、上机将上述配置好的体系放入荧光定量 PCR 仪中,激发和发射波长分别调至荧光染料相应波长, 升温速率通常设为 1 ℃/10s,温度变化范围一般从 25 ℃ 到 75 ℃ 或 100 ℃ 不等,每 1~3 ℃ 测定一次荧光强度(实验条件会根据实际情况优化)。5、数据分析通过数据处理软件,绘制出荧光强度随温度的变化曲线,通过拟合波尔兹曼(Boltzmann transport equation,BTE)方程,计算出蛋白样品的 Tm 值及相关参数。03 DSF优点1、Nano DSF无需染料:直接检测蛋白紫外荧光;2、初步或者大量检测蛋白与化合物之间是否有结合力;3、低样品消耗量:每个样品仅需10 μL,最低浓度仅需0.5 μg/mL;4、数据精度高:CV<0.1% (Tm);5、温度变化范围广。04 DSF应用1、蛋白质稳定性评估大多数蛋白质对温度要求较高,评估其热稳定性对更好地发挥其功能非常重要。DSF不但可以高通量测定蛋白质的热稳定参数,还能评估突变、pH、离子强度等其他因素对热稳定性的影响,筛选出提高热稳定性的因素。图3. DSF检测不同 pH 对蛋白热稳定性的影响2、蛋白质结构研究蛋白质结晶是研究蛋白质结构的一个非常重要的手段,DSF可以高通量筛选蛋白质的结晶条件。3、蛋白质-配体相互作用研究配体与蛋白质结合一般会导致蛋白质的热稳定性改变,DSF可以通过测定蛋白质的热稳定性参数评估配体与蛋白质之间的相互作用。4、蛋白质稳定剂、抑制剂的高通量筛选 通过对化合物配体库进行高通量筛选,获得与蛋白样品相结合的候选化合物,评估化合物对蛋白稳定性的影响,从而筛选出蛋白质稳定剂、抑制剂、辅助因子及分子伴侣。图4. DSF检测蛋白与配体的结合5、小分子化合物抑制机制研究 利用 DSF 可以进行高通量、矩阵式的多种化合物检测(化合物与一些已知的抑制剂、底物、产物或辅助因子等结合),用于小分子化合物抑制剂的筛选及其药理机制的研究。发布于 2022-11-10 17:21・IP 属地河南硕博材料表征​赞同 6​​添加评论​分享​喜欢​收藏​申请

基本完全指南:关于DSD DIFF DSF DST,播放和管理 - 知乎

基本完全指南:关于DSD DIFF DSF DST,播放和管理 - 知乎首发于折腾媒体库切换模式写文章登录/注册基本完全指南:关于DSD DIFF DSF DST,播放和管理How度You毒本人从不删评简单介绍几个名词,熟悉的请跳过DSD即(Direct Stream Digital)『直接比特流数字』,它是Sony与Philips在1996年宣布共同发展的高解析数字音响规格, DSD新技术与DVD的音响技术指针竞争,用1bit比特流的方式取样,采样率2.8224MHz [1] (CD 44.1kHz取样的64倍)的高取样方式,直接把模拟音乐讯号波形以脉冲方式转变为数字讯号,以将近四倍于CD的空间,储存音乐,因此可以提供更为优秀的声音效果,由于取样次数高,所以取样过的波形很圆顺,比较接近原来的模拟波形。——摘自百度百科简言之:DSD以1bit采样和记录,而传统采样和记录都采用多位bit,这是最大的不同,需要专门的解码器。DSD有着远高于CD的采样频率,更接近原始声音的波形。SACD即(Super Audio CD),可以简单理解为一种音乐碟片,以DSD方式记录声音信息。网上经常流传的是SACD iso文件(即SACD碟片的镜像文件)。DSD音乐原来是通过SACD(SuperCD,超级CD)来发行的,只有专门的DSD播放器(因为有专门的解码器)能播放,后来为了兼容普通的CD,就在一张碟上混合了普通的CD数据,叫做Hybird SACD(DSD和CD二合一混合光碟),这样,当插入DSD的播放器时,播放DSD音乐,插入普通的CD时,播放普通CD音质的音乐。SACD-R随着索尼的PS3被破解,导致SACD被刻盘出来了,即SACD Rip, 这样盗版出来的光碟叫做SACD-R。但这种盗版一般跟原版音质无差别,只是侵犯了别人的商业利益。DIFF 及 DSF只是两种将DSD封装后的(文件)格式,就像一篇文字,你可以封装(存放)在txt或doc格式的文件里。DIFF(即DSDDIFF)是飞利浦公司(Philips)推出的,DSF则是索尼公司(Sony)提出的。DIFF文件对tag信息支持不是太友好,DSF则很友好。关于tag管理,后文还有补充。DSTDSD文件很大,按原始方式存储很占空间,DST是DSD一种无损压缩方法。压缩文件音质不变,但播放时由于要“解压”,所以会增加不少运算量,有可能导致播放卡顿。SACD和DIFF都支持DST压缩,DSF尚不支持。无良DSD正统的DSD,应该在录音阶段就用DSD格式进行记录。很多“无良”产商,把1996年以前的录音带拿去用DSD转码器转成DSD后就拿出来卖,这种音乐的质量良莠不齐,但多少有些可取之处。还有一种更无良的做法,把普通CD升频转换为DSD,然后在换个包装,忽悠无知买家。这么做的一般都是盗版商,甚至是“网友出版”。(多说一句,目前网上很多的“网友出版”是将普通的两声道CD转换为DTS,所以遇到DTS最好睁大眼睛)播放DSD需要对应的解码器才能播放,这个解码器可以是硬件(如芯片),也可以是一个软件(如exe,dll,apk),它们解码出来的内容没有本质区别。DSD碟机,硬件解码,价格基本不亲民,而且歌曲放在碟片上管理也不大方便(相对于电脑tag管理而言)。手机和平板,安装nPlayer(nPlayer可能不是最好,但是日常使用真的很方便)。电脑上,Potplayer,Kodi(XMBC),AIMP(应该可以), JRiver等都能直接播放。foobar2000则需要加插件,大致有下边几种套路:懒人套路:foobar2000加FFmpeg Decoder Wrapper插件,然后在设置界面中添加.diff及.dsf两种格式即可,这种方式本质上是使用FFmpeg来解码。如果手头没有多好的音响设备,建议直接用这个方法。推荐套路:foobar2000加foo_input_sacd插件(“官网”:https://sourceforge.net/projects/sacddecoder/files),如果手头的音响支持DSD硬解码,可以再加装WASAPI或ASIO插件以发挥音响的全部能力。(WASAPI或ASIO都只是提供一个将音乐文件直接发送到音响的通道而已,使用该通道,本质是使用音响本身的硬件解码器)。不推荐:foobar2000加DSDIFF Decoder插件,一棍打死,不解释。 有网友评论留言想要解释: 这是我的个人选择,foo_input_sacd功能更多:DSDIFF Decoder插件功能单一,只能解码DIFF, 不支持DSF,不支持iso,不支持写入tag。还可能与另一个foo_input_sacd插件有冲突,而foo_input_sacd插件功能却全面得多,前述功能它全都支持。 但现在已经快2020年了,foo_input_sacd至今都不在foobar2000的官网插件列表里,DSDIFF Decoder插件却很长时间都在(现在也移除了),说明 Peter(foobar2000作者)大概还是更认可DSDIFF Decoder的。其实如果你经常使用foo_input_sacd,就会发现它有些缺陷(后文有提及),我猜这就是它至今没能“转正”的原因。 “鞋合不合脚,穿上就知道”:这些插件都是免费且公开的,一个才几MB,随手下载下来试用一下很快就知道谁更合适,比在这里打字留言快多了,我基本都不看评论,不看知乎通知(太多看不过来),也不关心文章的阅读数据,毕竟不是靠这些恰饭的,,当然看到有人点赞还是比较开心的。 格式转换从SACD iso提取DIFF或DSF用sacd_extract.exe,这是命令行工具,官方下载(https://github.com/sacd-ripper/sacd-ripper/releases)另外还有一个对应的sacd_extract GUI.exe 这是GUI界面,方便普通用户使用。从DIFF转换为DSF用dff2dsf.exe,这是一个命令行工具。另外还有一个对应的DFF2DSF-Shell.exe,这是一个GUI界面工具,方便普通用户使用。从DSF转换为DIFF尚未找到直接转换工具,如有发现,欢迎留言。在DIFF上应用DST压缩找Philips_ProTECH_DST_Encoder,请注意目前只支持DIFF格式压缩,压缩后文件体积变小,扩展名仍为diff。DIFF,DSF转换为SACD找Philips_SuperAuthor。DIFF里的 DST解压没找到直接转换工具,估计先制作成SACD然后借助sacd_extract.exe在提取轨道时可以选择顺便解压。不过这难度和繁琐程度估计可以劝退不少人。顺便说一下这些网上流出的转换工具,它们多半是破解索尼或飞利浦的设备所得,只有二进制可执行文件,没有源代码,用法并不灵活。目前我都尽量保留原始的iso文件,因为iso转其他格式都很方便,而其他格式的转换多不太方便。tag管理(标签信息管理)很多SACD iso文件,包括diff或dsf文件上的tag信息都是一堆英语或拼音,甚至是TRACK01,TRACK02这样无意义的文字,查阅很不方便,对于中文歌曲,我们还是希望能以中文表示歌曲名,歌手等信息。那么该如何做呢?以foobar2000加foo_input_sacd插件为例安装完该插件后,在其设置界面上勾选“Editable Tags”和“Store Tags With ISO”两项后,点确定。然后: 对于iso文件:直接在foobar里修改tag信息即可,这不会改变iso文件本身,而是相同目录下生成一个对应的xml文件。 对于diff文件:直接在foobar里修改tag信息即可,这会改变diff文件本身,foobar2000目前版本(1.3.17)对diff的tag读取并不是很好,有时会读不出修改后的信息。题外话,我怀疑这不是foobar播放器本身的Bug,因为我通过二进制分析,发现tag写入的位置比较混乱,而且有些改了有些没改。 对于dsf文件:foobar目前可以完美读取tag,但无法修改,不过可以通过一款名为mp3tag软件来修改,该软件的官网(https://www.mp3tag.de/en/download.html)。 ——2018-05-14——2019-11-19,重新整理文章格式排版,并添加了若干评论区关心的内容。编辑于 2019-11-23 16:21Foobar2000高保真音频(High Fidelity, Hi-Fi)DSD​赞同 174​​27 条评论​分享​喜欢​收藏​申请转载​文章被以下专栏收录折腾媒体库介绍各种音乐/视频/图片等多媒体软件硬件(个

Object reference not set to an instance of an object.

Object reference not set to an instance of an object.

Server Error in '/' Application.

Object reference not set to an instance of an object.

Description: An unhandled exception occurred during the execution of the current web request. Please review the stack trace for more information about the error and where it originated in the code.

Exception Details: System.NullReferenceException: Object reference not set to an instance of an object.

Source Error:

Line 28: if (Request["lang"] == null && Session["lang"] == null)

Line 29: {

Line 30: Session["lang"] = Request.UserLanguages[0];

Line 31: }

Line 32: else if (Request["lang"] != null)

Source File: e:\InetPub\dsfhome.1\App_Code\basePage.cs    Line: 30

Stack Trace:

[NullReferenceException: Object reference not set to an instance of an object.]

Localization.basePage.InitializeCulture() in e:\InetPub\dsfhome.1\App_Code\basePage.cs:30

ASP.default_aspx.__BuildControlTree(default_aspx __ctrl) in e:\InetPub\dsfhome.1\Default.aspx:1

ASP.default_aspx.FrameworkInitialize() in e:\InetPub\dsfhome.1\Default.aspx.cs:912308

System.Web.UI.Page.ProcessRequest(Boolean includeStagesBeforeAsyncPoint, Boolean includeStagesAfterAsyncPoint) +56

System.Web.UI.Page.ProcessRequest() +80

System.Web.UI.Page.ProcessRequestWithNoAssert(HttpContext context) +21

System.Web.UI.Page.ProcessRequest(HttpContext context) +49

ASP.default_aspx.ProcessRequest(HttpContext context) in c:\Windows\Microsoft.NET\Framework\v2.0.50727\Temporary ASP.NET Files\root\fb509fc4\c07da73a\App_Web_duy7nmsk.1.cs:0

System.Web.CallHandlerExecutionStep.System.Web.HttpApplication.IExecutionStep.Execute() +181

System.Web.HttpApplication.ExecuteStep(IExecutionStep step, Boolean& completedSynchronously) +75

Version Information: Microsoft .NET Framework Version:2.0.50727.8745; ASP.NET Version:2.0.50727.8745

如何播放.dsf文件? | Sony China

如何播放.dsf文件? | Sony China

跳至內容

文章ID : S800009143 / 最近修改 : 2018-02-22打印如何播放.dsf文件?

该文章适用的产品及类别如何播放.dsf文件?DSD格式的音频文件(.dsf)包含有高质量的DSD音频数据,可以使用装有DSD播放插件的Windows Media Player来播放。请接受 Youtube Cookie 以观看此视频在下面访问您的 Cookie 首选项,并确保在“功能”部分下打开 Youtube Cookie。管理 Cookie

相关文章关于Linux机型如何播放本地多媒体文件?关于Linux机型如何在浏览器中播放视频/音频文件?在Windows XP中如何使用备份工具备份文件和文件夹?如何使系统支持更多的文件类型?如何使用WinDVD/WinDVD BD播放影片? 如何以全屏模式播放影片?

Differential Scanning Fluorimetry (DSF) | Center for Macromolecular Interactions

Differential Scanning Fluorimetry (DSF) | Center for Macromolecular Interactions

Skip to main content

Main MenuUtility MenuSearch

HARVARD.EDU

 

Enter PPMS

Search

HomeAboutTeamDirections to the CMINews & EventsCMI SupportContactAccess & FeesPoliciesCMI Lab SafetyFeesCMI Account CreationFormsGetting Started at the CMITechnologiesInstrumentation OverviewSurface Plasmon ResonanceBiolayer InterferometryMicroScale ThermophoresisIsothermal Titration CalorimetryCircular DichroismDifferential Scanning FluorimetryLight ScatteringDynamic Light ScatteringSEC-MALSMass PhotometryServicesTrainingData Collection ServicesLight Scattering ServicesDSF ServicesCD ServicesProtein Purification ServicesData Collection Service FeesNanobody ServicesResourcesUser AgreementsProtein ResourcesMeasuring ConcentrationCMI Instrument ResourcesGrant SupportData ManagementAssociated Core FacilitiesUser PublicationsCMI Knowledge BasePublications

HOME / TECHNOLOGIES /

Differential Scanning Fluorimetry (DSF)

For information on access fees, policies and getting started at the CMI, see the CMI Access Page.

DSF at the CMI

Differential Scanning Fluorimetry (DSF) measures protein unfolding by monitory changes in fluorescence as a function of temperature. Conventional DSF uses a hydrophobic fluorescent dye that binds to proteins as they unfold. NanoDSF measures changes in intrinsic protein fluorescence as proteins unfold.

The CMI has a modified Life Technologies Quant Studio 6/7, for conventional DSF. The CMI has the Life Technologies Protein Thermal Shift Analysis software for DSF data fitting.

The CMI has a Prometheus NT.Plex instrument from NanoTemper Technologies with aggregation optics. The CMI has these Data collection and analysis software packages: PR.ThermControl for thermal stability data collection, PR.ChemControl for chemical stability data collection, PR.TimeControl for time interval data collection, and PR.Stability Analysis for advanced data analysis.

 

DSF (Conventional DSF, Protein Thermal Shift Analysis)

Conventional Differential Scanning Fluorimetry (DSF) uses a real-time PCR instrument to monitor thermally induced protein denaturation by measuring changes in fluorescence of a dye that binds preferentially to unfolded protein (such as Sypro Orange, which binds to hydrophobic regions of proteins exposed by unfolding).  This experiment is also known as a Protein Thermal Shift Assay, because shifts in the apparent melting temperature can be measured upon the addition of stabilizing or destabilizing binding partners or buffer components.

NanoDSF

NanoDSF is a modified differential scanning fluorimetry method which monitors intrinsic tryptophan and tyrosine fluorescence as a function of temperature, time, or denaturant concentration. Tryptophan and tyrosine fluorescence intensity and wavelength maximum will vary as the local chemical environment changes, with significant changes occurring as buried or packed aromatic side chains become solvent exposed upon unfolding.  NanoDSF measures fluorescence intensity at 350 nm and 330 nm and compares the ratio as a function of temperature or denaturant concentration. NanoDSF can be used for a broader range of protein samples than traditional DSF and has significantly higher throughput and lower sample consumption than DSC or CD. Free energies of folding and temperatures of unfolding measured using NanoDSF are comparable to values determined by DSC for a range of sample types. The only significant limitation is that the protein of interest must contain aromatic amino acids (tryptophan or tyrosine).

 

Data Files - About CMI Data Files

Users are responsible for storage of all raw and processed data collected at the CMI.

Users should have a plan to copy or transfer all raw and process data to their own local or cloud storage system.

While the CMI allows temporary local storage of CMI User data on the instrument computer, we make no guarantees on the security or long-term availability of any data at the CMI.

For most (but not all) CMI technologies, the raw data files and recommended readable exports are relatively small and can be readily transferred electronically. 

See specific instruments for exceptions and for details about the software, data file types and recommended data exports. 

Data Sharing:

Currently, a Generalist Repository is the recommended data repository for most CMI data types, as stable specialist data repositories have not been established.

Data Files - DSF - QuantStudio 6/7

Technology

Differential Scanning Fluorimetry (DSF)

Instrument

Life Technologies Quant Studio 6/7

Recommended Repository

Generalist Repository

 

 

 

 

 

 

Software Type

Data Collection

Current Version

QS Real-Time PCR Software, version 1.7.1

Data Files (Type, ~size)

experiment file

.eds

2-10 MB/plate

 

 

 

 

Software Type

Data Analysis

Current Version

Applied Biosystems Protein Thermal Shift, version 1.2

Data Files (Type, ~size)

experiment file

.eds

2-10 MB/plate

Readable Exports

raw data

.csv

2 MB/plate

 

analyzed data

.csv

12 KB/project

 

analyzed data

.txt

29 KB/project

Data Files - DSF - Prometheus

Technology

Nano Differential Scanning Fluorimetry (nanoDSF)

Instrument

NanoTemper Prometheus NT.Plex

Recommended Repository

Generalist Repository

 

 

 

 

Software Type

Data Collection (Thermal Denaturation)

Current Version

PR.ThermControl, Version 2.3.1

Data Files (Type, ~size)

experiment file

.prc

10-30 MB/project

 

raw data

.xslx

2 MB/project

 

 

 

 

Software Type

Data Analysis

Current Version

PR.Stability Analysis, Version 1.1

Data Files (Type, ~size)

analysis file

.pra

2-6 MB/project

Readable Exports

processed data

.xslx

~500 KB/sample

 

results table

.xslx

~30 KB/project

 

 

 

 

Software Type

Data Collection (Chemical Denaturation)

Current Version

PR.ChemControl, Version 1.4.3

Data Files (Type, ~size)

experiment file

.prcc

10 MB/project

Readable Exports

raw data

.xslx

6 KB/sample

 

 

 

 

Software Type

Data Collection (Time Control)

Current Version

PR.TimeControl, Version 1.0.2

Data Files (Type, ~size)

experiment file

.prtime

2 MB/project

Readable Exports

raw data

.xslx

60 KB/project

DSF Data Collection Services

DSF Service Overview

In addition to instrument training, the CMI is now offering basic protein Differential Scanning Fluorimetry services, including protein thermal shift analysis and buffer optimization.

Differential Scanning Fluorimetry (DSF) with fluorescent dye or using intrinsic protein fluorescence

Life Technologies Quant Studio 6/7

Conventional DSF, protein thermal stability using hydrophobic dye Sypro Orange

Buffer optimization

NanoTemper Technologies Prometheus NT.Plex

NanoDSF, protein thermal stability using intrinsic protein fluorescence

Chemical denaturation

Buffer optimization

Data Collection Fees Summary

Data Collection

Limited Data Collection Services are offered.

Service fees are based on labor and supplies costs, and will be charged for all completed services, regardless of experimental outcome. 

Before submitting samples for data collection, users must approve the estimated charges and be given a date and time for sample delivery.

External Users will also be required to submit a PO and a signed CMI User Agreement.

Most CMI Data Collection Services include a setup fee plus a per-sample data collection fee.

Some services include replicate measurements by default in the per-sample fee. For others, there is a reduced-price replicate measurement fee, if collected in the same dataset.

Nanobody services not available to commercial users at this time.

Current Harvard Life Lab commercial users are offered a 25% discount off the standard commercial rates.

All Data Collection Fees

DSF Resources (Conventional DSF)Quant Studio DSF SuppliesQuant Studio qPCR Resources

 

CMI QuantStudio DSF Getting Started Guide

Protein Thermal Shift Studies manual from Applied Biosystems by Life Technologies

All Experiments:96-well FAST-block optical plate, eg.: LifeTechnologies MicroAmp FAST optical 96-well reaction plate, 0.1 ml, 4346907optical adhesive film, eg.: LifeTechnologies MicroAmp Optical Adhesive Film, 4360954DSF/Protein Thermal Shift ExperimentsDSF compatible dye, eg.: LifeTechnologies Protein Thermal Shift Dye Kit, 4461146 (Sypro Orange)samples, ligands, buffers qPCR ExperimentsqPCR reagents (eg. LifeTechnologies PowerUp SYBR Green Master Mix, A25742)primers and templates 

qPCR Resources

CMI QuantStudio qPCR Getting Started Guide

Quant Studio 6/7 Quick Reference Guide from Applied Biosystems by Life Technologies

NanoDSF ResourcesPrometheus NT.Plex Supplies

NanoDSF Resources

CMI Prometheus NanoDSF Getting Started Guide

 

NanoDSF Technology Page NanoTemper Technologies

NanoDSF Supplies

NT.Plex Capillary Chips

2x 8 Standard 24-Capillary Chips, NanoTemper Catalog # PR-AC002

2x 8 High Sensitivity 24-Capillary Chips, NanoTemper Catalog # PR-AC006

384-well plates for loading capillaries

Protein samples, ligands, buffers

See also: molecular properties, Differential Scanning Fluorimetry, DSF

TechnologiesInstrumentation OverviewSurface Plasmon Resonance (SPR)Biolayer Interferometry (BLI)MicroScale Thermophoresis (MST)Isothermal Titration Calorimetry (ITC)Circular Dichroism (CD)Differential Scanning Fluorimetry (DSF)Light Scattering

Center for Macromolecular InteractionsHarvard Medical School240 Longwood Ave.Building C, Room C-303Boston, MA 02115

cmi@hms.harvard.edu(617) 432-5004

Harvard Medical SchoolBiological Chemistry and Molecular PharmacologyHMS Core Facilities

8e07e3599f155c7eaba3c22aae9f5b63

Harvard University Privacy Statement

Admin Login

Copyright © 2024 The President and Fellows of Harvard College | Privacy | Accessibility | Digital Accessibility | Report Copyright Infringement

dsf职业年金什么意思啊? - 知乎

dsf职业年金什么意思啊? - 知乎首页知乎知学堂发现等你来答​切换模式登录/注册养老保险企业年金保险投资万能险年金险dsf职业年金什么意思啊?关注者5被浏览2,079关注问题​写回答​邀请回答​好问题​添加评论​分享​2 个回答默认排序麦尼财经​CFA 特许金融分析师资格证持证人​ 关注关于社保,企业年金,个人养老金账户,很多人都分不清楚,只知道是和养老有关。前面已经有文章详细讲解了社保(参考旧文《退休后,社保能领多少钱》),和个人养老金账户(参考旧文《个人养老金账户》),这里就企业年金做一个详细说明,建议收藏+关注,就算你现在没有企业年金,说不定过几年就有了呢。01 什么是企业年金?企业年金是公司为员工养老提供的一种福利计划。我们都知道,现在一般公司都会给员工上社保,给员工提供一个基本保障。有条件的公司,除了社保之外,还给员工多交一部分钱,也就是企业年金,经常听到有人说的五险二金,或者六险二金。这“二金”就是住房公积金和企业年金。和社保不同的是,社保比较普遍,企业年金只有一些大公司,或者盈利状况比较好的公司才有。因为老龄化越来越严重,政府也鼓励企业设立企业年金账户,给员工做养老补充,所以近几年,也有越来越多的企业开始设立企业年金。和企业年金类似的,还有职业年金。职业年金是机关事业单位的养老补充计划,其性质及运营模式和企业年金基本一致。02 哪些人享有?企业年金根据人社部披露的数据,截止2022年底,共有12.8万家企业设有企业年金,参加的职工共有3010.29万。对于个人来说,能否享受企业年金的福利,完全取决于自己所在的公司是否设有企业年金。企业年金对于企业来说,是一笔不小的成本,但是也有越来越多的企业设立企业年金,以此来吸引人才并且留住员工。职业年金职业年金是强制性的,不是单位想设立就设立,不想设立就不设立的。所有的机关事业单位都有职业年金,比如医生、老师、公务员,都享有职业年金。03 现在交多少钱?谁来交钱?企业年金企业年金在缴费上,和社保类似,都是由企业和员工共同承担。按照最初提出企业年金时候的规定:双方缴费金额合计不能超过员工工资的12%。其中企业缴费不能超过8%,各自缴费的比例,也可以由企业和员工共同协商决定。但是现在,这个比例也有所调整,现在企业缴纳的比例是不设上限的。职业年金职业年金是强制性的,其缴费标准也是强制性的,没有什么弹性空间。就是单位缴纳员工工资的8%,个人缴纳4%。个人缴纳部分通过企业代扣代缴,也就是发工资之前,先把这一部分扣除了,从个人的税前收入里扣,扣除完这部分之后的工资所得,再扣个税。企业(单位)缴纳部分,8%以下部分,也是从税前扣,也就是给员工缴纳企业年金(职业年金)的这部分钱,是不用交税的。8%以上部分,需要从企业税后所得里扣除。所以,一般情况下,企业缴纳金额都在8%。总的来说,就是工资越高,企业年金越多。04 这些钱谁来管理?很多人都担心:如果企业活不到我退休呢?就算公司有企业年金,但是还没等到我退休,公司就破产了,那企业年金还有用吗?我退休以后还能领到钱吗?这个问题不用担心,企业年金的钱是专门设立账户的,这些钱由专门的资管机构做投资管理。截止2022年底,全国企业年金累计总规模为2.87万亿元。不同的企业,企业年金可能会托管在不同的资管机构:看到这,可能又有一个问题:这些资管机构管理,会不会亏损呢?客观来说,所有的投资都有可能亏损。企业年金也不例外,比如去年,也就是2022年,企业年金收益率为-1.83%。企业年金的投资管理风险还是相对比较低的,其投资标的和投资范围都有严格的限制,大部分比例都是投资风险等级低的金融产品,所以长期看来,不太可能出现大幅亏损,偶尔出现小幅亏损也在所难免。05 中途离职,这些钱还能拿到吗?这也是很高频的问题,如果离职了,企业年金还能拿到吗?企业年金的钱分为两部分,公司交的和自己交的,分别在企业账户和个人账户。个人账户的钱,一直都是自己的,即使离职了,这部分也在自己的账户里。等到退休的时候,可以领取个人账户里的钱。企业账户里的钱到底能不能领取,这个比较复杂,具体要看企业的规定。比如有的企业规定,要工作满5年,才兑现一年的企业年金,5年以后,每多一年,就多兑现一年的钱。如果5年之内离职,就一点也享受不到企业缴纳的那部分钱。也有比较人性化的企业,不管工作几年,企业缴纳部分全部转给员工让其带走。但是为了留住员工,一般企业都有年限限制。当然,企业设置这个年限,也受约束,最长不能超过8年。当然,也有的企业,没等你离职,也没等你退休,因为效益不好,就先暂停企业年金了。06 什么时候开始领取?最后,就是大家最关心的问题,企业交的这笔钱,什么时候能领?从一开始就说了,这就是个补充养老的福利,所以得等到退休养老的时候才能领。退休之后,有两种领取方式:可以一次性领出来,也可以按月领取。企业账户和个人账户的钱是随时能查看的,所以退休的时候,能看到一共有多少钱。一次性领取就是账户当时的累计金额一笔领出。按月领取的话,先计算出领取月数,可能按照企业预估的员工平均寿命,退休之后还剩余多少个月,就把账户累计金额均摊到这些月发放;也可能按照企业约定的固定年限,比如发放二十年,就把账户累计金额均摊到这二十年的每个月发放。在领取上,和社保有两点不同:第一,社保是交满15年才能领取,企业年金没有这个强制要求,对这个年限要求比较灵活,各企业规定都不一样。第二,社保养老金是领取至终身,企业年金是把账户里的钱领完为止。什么时候领完,就停止不再发放了。最后,还有一个问题,如果企业年金账户里的钱还没领完,人就不在了,那剩余的钱怎么办呢?账户里剩余的金额是可以继承的,如果是这种情况,剩余的钱就作为遗产交给继承人。发布于 2023-04-17 22:04​赞同​​添加评论​分享​收藏​喜欢收起​牛肉西红柿交通银行客户经理​ 关注d+sf养老金是dsf职业年金哦!指事业单位和行政单位为雇员或公职人员提供补充养老保险的总称。在国家统一监督、指导下,由公职人员和事业单位共同缴费组成养老保险费。为公职人员提供退休收入保障的养老金保险制度。亲,dsf养老金是每个月发一次,而不是一年一次。我国的职业年金是社保之外的另一个补充xing养老保险,在职时按月按比例交费,由单位和个人共同承担,建立个人的专门帐户,办理退休后按月按照计发月数以固定金额的形式予以发放。按单位8%和个人4%存入养老保险个人账户。退休后有两种方式领取1.是全部购买商业保险,按约定每月领取。2.是按退休法对应的计发月数发,发完为止。社保局还会对某些自主创业或者灵活就业人员进行鼓励性补贴。发布于 2023-04-14 13:02​赞同​​添加评论​分享​收藏​喜欢收起​​

上海交大何亚文团队联合全球相关科研人员发表“群体感应信号DSF介导的种内、种间和跨界信号交流”综述文章_交大智慧_上海交通大学新闻学术网

上海交大何亚文团队联合全球相关科研人员发表“群体感应信号DSF介导的种内、种间和跨界信号交流”综述文章_交大智慧_上海交通大学新闻学术网

探索发现

首页

探索发现

交大智慧

正文

探索发现 · 交大智慧

上海交大何亚文团队联合全球相关科研人员发表“群体感应信号DSF介导的种内、种间和跨界信号交流”综述文章

2022年09月13日

责任编辑:冯硕 

近日,国际微生物学权威综述期刊Trends in Microbiology在线发表了上海交通大学生命科学技术学院何亚文教授团队的综述文章“DSF家族群体感应信号介导的种内、种间和跨界信号交流 (DSF-family quorum sensing signal-mediated intraspecies, interspecies, and inter-kingdom communication)”。何亚文教授为综述第一作者和通讯作者,中山大学邓音乐教授、新加坡南洋理工大学缪岩松教授(Prof. Yansong Miao)为综述共同第一作者,西北大学田静教授、印度DNA指纹识别和诊断中心(Centre for DNA Fingerprinting & Diagnostics) Subhadeep Chatterjee研究员、美国加州大学伯克利分校(University of California, Berkeley)Steven Lindow院士和南阿拉巴马(University of South Alabama) Tuan Minh Tran助理教授参与了该综述文章的写作。DSF (Diffusible signaling factor)信号分子是多种革兰氏阴性细菌产生的一类中链顺式不饱和脂肪酸(图1)。这些微生物通过识别自身合成与分泌的DSF,感应自身群体细胞密度,诱导相关基因的表达,相应调整细胞代谢与生理状态。根据信号接收与传导机制,DSF介导的种内交流主要分为三类:(I)植物病原黄单胞菌为代表,DSF由RpfC/RpfG双组分系统接收与传导。上海交通大学何亚文教授团队一直引领这一研究领域,先后鉴定了黄单胞菌中DSF家族信号分子的化学结构、信号传导途径和调控的生物学功能,阐明了DSF生物合成途径与DSF翻转(turnover)的分子机制, 进一步发现寄主植物免疫系统与黄单胞菌DSF群体感应系统之间存在着精细的相互作用(图2)。(II)动植物病原伯克氏菌为代表,DSF由调控蛋白RpfR接收与传导。中山大学邓音乐教授团队一直引领这一研究领域(图3)。(III)以条件致病绿脓杆菌为代表,DSF接收与传导机理尚待研究。图1 三种DSF-介导的种内交流系统及其所依赖的信号分子图2 野油菜黄单胞菌中DSF生物合成与信号通路图3 伯克氏菌感应BDSF信号及其传导通路DSF产生菌也能够利用DSF信号与其它细菌之间开展种间交流(图4),这些细菌包括伯克氏菌(Burkholderia spp.)、嗜麦芽寡养单胞菌(Stenotrophomonas maltophilia)、绿脓杆菌(Pseudomonas aeruginosa)、弗郎西斯菌(Francisella novicida)、芽孢杆菌(Bacillus)、噬菌蛭弧菌(Bdellovibrio bacteriovorus)、沙门氏菌(Salmonella)等。在目前发表的文献中,多数DSF产生菌均利用分泌的DSF信号分子干扰其它细菌的生长或发育,限制其发展,保持自身在局部环境的竞争力。DSF-产生菌还能与真核生物之间开展跨界信号交流(图5)。DSF家族信号分子显著影响白色念珠菌(Candida albicans)菌丝体发育和生物膜形成,影响其致病性。加州大学伯克利分校Steven Lindow教授团队系统阐明了苛养木杆菌利用DSF信号分子调控其在利器叶蝉(sharp shooter leaf hoppers)前肠中的定殖能力,有利于苛养木杆菌通过这种刺吸式昆虫转移到葡萄等寄主植物维管束中。南洋理工大学缪岩松教授团队首先发现DSF可以诱导植物细胞膜甾醇富集,高度诱导植物细胞壁纤维素合成,以及影响细胞膜上免疫功能分子的生化特性,比如植物鞭毛蛋白受体聚集和内吞,从而抑制植物病程相关分子模式激发的免疫反应(PTI);进一步研究发现DSF 影响植物细胞壁-质膜-微丝骨架复合体[plant cell wall-plasma membrane-actin cytoskeleton (CW-PM-AC) continuum]的生物物理特性, 扰乱细胞膜表面上蛋白分子的识别与大分子组装,导致植物免疫系统信号失调 (图5)。西北大学田静教授团队发现DSF信号可以有效抑制斑马鱼中脂多糖(LPS)诱导的炎症反应,其作用机理可能是干扰Toll-样受体通路抑制下游炎症因子的表达以及影响caspase 家族级联反应激活,抑制溶酶体组织蛋白酶基因表达从而减轻炎症导致的细胞凋亡。阐明DSF介导的种内、种间与跨界交流现象与机制进一步深化了微生物社会学的内涵,同时也有助于研发新型抗感染策略图4 DSF-介导的种间与跨界信号交流图5 DSF在拟南芥和斑马鱼中的信号途径本研究得到了国家重点研发计划、国家自然科学基金和上海农乐生物制品股份有限公司资助。感谢张炼辉(Lian-Hui Zhang)教授的指导、支持与鼓励。论文链接:https://www.cell.com/action/showPdf?pii=S0966-842X%2822%2900188-3

作者:

生命科学技术学院

供稿单位:

生命科学技术学院

打印本页

关闭窗口

DSF是什么? - 知乎

DSF是什么? - 知乎首发于DSF——全球首个DeFi社交平台切换模式写文章登录/注册DSF是什么?得道社区 DSF是全球首个已经落地使用的去中心化社交金融区块链平台,也是革新现有MakerDAO和DeFi,以及承载互联网金融行业转型的项目中,公认的最受瞩目最为可行的实现方案。 DSF网络具有自主扩展和极速裂变的特点。在DSF网络中,每个节点用户都可向下开发,构建起一个巨大的无穷尽的社交网络,从而有能力去承载丰富多样的区块链金融产品。 相对于现有区块链金融平台,DSF为链上社交金融垂直领域带来了一个极高可扩展性,且极低成本的分布式应用运行、开发、维护和裂变平台,给诸多挣扎在高昂开发成本,技术难关和流量封锁的区块链金融产品团队,或正在艰难转型的互联网金融企业带来曙光。 DeChat则是DSF的首个落地杀手级DAPP,携1200万用户城池部署上链,正是由于DeChat如此超凡的竞争力,不少人认为DSF很有可能在2020成为巨无霸项目。DSF正是DeChat平台通证DE的映射生态母币。 除此之外,从最近的情况来分析,DSF链上社交金融平台开发中最困难的环节已经接近完成,在5月发布的Q3 Roadmap中,我们可以看到DSF的三大重要更新,体现了项目核心团队过去两年对区块链技术落地演进的高度思考,分别是: 1)去中心化社交网络,具备无限扩展和激励机制的社交关系上链; 2)去中心化金融钱包,通过智能合约调用用户社交网络,并提供社交金融类服务; 3)不同类型的智能合约,用于社交关系确认,基于社交关系的系统挖矿奖励,以及区块链金融产品对社交关系的使用。 正如IPFS的出现颠覆了分布式储存和共享文件的网络传输逻辑,我们认为DSF的三大核心功能更新也将改变“链上·社交·金融”现有的运转轨迹: 社交关系上链→形成去中心化社交网络→通过智能合约调度去中心化社交网络服务于区块链/互联网金融产品的开发、运营、传播→DSF Token全流程、全生态嵌入/治理/激励→社交推进金融,金融促进社交。 至此,一个更安全、便捷、极具滚雪球效应的,可无穷延展的链上社交金融生态网络闭环形成。风口之下,DSF Token的价值预期在于其顶层商业设计、资本实力、用户城池壁垒,以及团队在垂直领域的多年深耕,要知道设计出一款具有变革意义,并且自传播裂变效果很好的产品绝非易事,这是团队在社交、金融、区块链行业经验、渠道、资源质变的结果。 DSF就像是一把开启链上社交金融2.0时代新大门的钥匙,正以我们看不见的速度飞快发展,为行业打开全新的局面,越来越丰富的产品矩阵也会随之出现。 届时,DSF Token的流动性,DeChat日均UV和社交网络的裂变效应也将愈发明显,而这些最终都将表现在其通证的筑底价格稳固和币价上涨速度上。 一句话总结,DSF有打造出一个超量级去中心化社交金融网络的势能,有落地产品,有钱包入口,有盈利模型,有用户基础,并处于爆发增长赛道。 总而言之,无论是区块链社交还是Defi,DeChat团队都展现出了其强大的底蕴,特别是1200万用户这一必杀技,更是秒杀了行业内同类型的项目,使那些项目瞬间相形见绌,根本无法与DeChat相提并论。 实际上,行业内的反响也正是这样,无论是DeChat在朋友圈的火爆,还是ZB交易所的提前预定,都从侧面证明了DeChat的实力。 2020年转眼已经过半,如果说今年真的会出现一个现象级的项目的话,我认为DeChat绝对有这样的实力。 如果在你的投资布局中,有一部分是准备配比给一些新兴潜力领域的,DSF应该在其中。2014年那波熊市中诞生的以太坊,不就是这么一路走过来的千倍币吗?发布于 2020-06-28 15:01去中心化区块链技术token​赞同 1​​1 条评论​分享​喜欢​收藏​申请转载​文章被以下专栏收录DSF——全球首个DeFi社

Theory and applications of differential scanning fluorimetry in early-stage drug discovery | Biophysical Reviews

Theory and applications of differential scanning fluorimetry in early-stage drug discovery | Biophysical Reviews

Skip to main content

Advertisement

Log in

Menu

Find a journal

Publish with us

Track your research

Search

Cart

Home

Biophysical Reviews

Article

Theory and applications of differential scanning fluorimetry in early-stage drug discovery

Review

Open access

Published: 31 January 2020

Volume 12, pages 85–104, (2020)

Cite this article

Download PDF

You have full access to this open access article

Biophysical Reviews

Aims and scope

Submit manuscript

Theory and applications of differential scanning fluorimetry in early-stage drug discovery

Download PDF

Kai Gao1, Rick Oerlemans1 & Matthew R. Groves 

ORCID: orcid.org/0000-0001-9859-51771 

40k Accesses

117 Citations

2 Altmetric

Explore all metrics

AbstractDifferential scanning fluorimetry (DSF) is an accessible, rapid, and economical biophysical technique that has seen many applications over the years, ranging from protein folding state detection to the identification of ligands that bind to the target protein. In this review, we discuss the theory, applications, and limitations of DSF, including the latest applications of DSF by ourselves and other researchers. We show that DSF is a powerful high-throughput tool in early drug discovery efforts. We place DSF in the context of other biophysical methods frequently used in drug discovery and highlight their benefits and downsides. We illustrate the uses of DSF in protein buffer optimization for stability, refolding, and crystallization purposes and provide several examples of each. We also show the use of DSF in a more downstream application, where it is used as an in vivo validation tool of ligand-target interaction in cell assays. Although DSF is a potent tool in buffer optimization and large chemical library screens when it comes to ligand-binding validation and optimization, orthogonal techniques are recommended as DSF is prone to false positives and negatives.

Similar content being viewed by others

Quantitative Mass Spectrometry-Based Proteomics: An Overview

Chapter

© 2021

Software for molecular docking: a review

Article

16 January 2017

Nataraj S. Pagadala, Khajamohiddin Syed & Jack Tuszynski

Dynamic light scattering: a practical guide and applications in biomedical sciences

Article

06 October 2016

Jörg Stetefeld, Sean A. McKenna & Trushar R. Patel

Use our pre-submission checklist

Avoid common mistakes on your manuscript.

IntroductionBiophysics drives modern drug discovery efforts, allowing rapid and high-throughput data acquisition to screen through large compound libraries in an effort to identify new bioactive molecules. An important component of this biophysics armory is the thermal shift assay, also commonly known as differential scanning fluorimetry (DSF) (Semisotnov et al. 1991). DSF is a cost-effective, parallelizable, practical, and accessible biophysical technique widely used as a method to track both protein folding state and thermal stability. It provides a reliable tool to examine protein unfolding by slowly heating it up in a controlled environment. By measuring the corresponding changes in fluorescence emission upon temperature increase, the process of protein denaturation can be monitored. Since changes in sample behavior through complex formation with even weakly binding ligands affect protein thermal stability, the technique has seen many successful applications and has been used in different ways over recent years. It has been utilized primarily as a drug discovery method to identify promising lead compounds for a number of target proteins for decades (Pantoliano et al. 2001). Another major application for DSF is in protein buffer optimization, identifying optimal conditions for storage, assay screening, and crystallization. By screening sparse matrix conditions, encompassing different buffer systems that cover a wide range of pH, additives, and salt concentrations, optimal buffer components can be identified for each individual protein. This has been shown to increase the success rates of protein crystallization in past decades (Huynh and Partch 2015). More recently, DSF has also been applied to the challenge of sample preparation, with two publications demonstrating that suitable screening approaches can be used to identify and optimize sample refolding buffers—allowing significantly cheaper access to the amounts of protein sample required to support high-throughput screening campaigns (Biter et al. 2016; Wang et al. 2017). Finally, a very recent development has shown that DSF is able to provide reliable data in complex solutions, such as unpurified chemical reactions. This is an exciting development, as the production and purification of chemical entities are a major bottleneck in any screening campaign.While the robustness of the DSF method and its broad applicability in both sample preparation and screening has led it to become an important biophysical tool in drug discovery, it is important to bear its limitations in mind. This is particularly true when designing a screening campaign, as such a campaign should contain orthogonal screening options that are not susceptible to similar limitations—in order to minimize both false positives and false negatives.In this review, we will provide a theoretical background of DSF as well as examples of its use in the various aspects of drug discovery introduced above—including the latest applications of DSF by ourselves and other researchers. We will also attempt to place DSF within the variety of biophysical methods currently used in screening campaigns and highlight areas of overlap or mutual limitations.Theory of differential scanning fluorimetryIn 1997, Pantoliano et al. (1997) introduced a new thermal shift assay system used in the screening of combinatorial libraries against different receptor proteins. Compared with conventional methods of the time, such as those based on calorimetry and spectral technologies (Bouvier and Wiley 1994; Weber et al. 1994), the newly developed system could implement high-throughput screening instead of assaying a single condition at a time. The custom-designed 96- or 384-well plates and fluorescence readout apparatus could easily monitor protein unfolding in multiple conditions, with different ligands and/or at different ligand concentrations in a single experiment. This helped researchers overcome a lot of cumbersome, slow, and labor-intensive work required by traditional methods. Rather than the need for a dedicated device, many labs already possess (or have access to) real-time polymerase chain reaction (RT-PCR) equipment that allows for fluorescence measurements over a controlled temperature range. Access to such equipment, the development of more sensitive dyes, and improved protocol design drove the use of DSF (Niesen et al. 2007).Proteins are in a thermodynamic equilibrium between folded and unfolded states (Bowling et al. 2016). An increase in energy of the environment (i.e., increase in temperature) pushes a protein toward the unfolded state which, when quantified, allows for the determination of the melting temperature (Tm), defined as the temperature at which 50% of a protein sample is in folded and 50% is in an unfolded state (Lo et al. 2004) (Fig. 1a). A change in the protein environment (including pH, ionic strength, or the presence of specific anions or cations) and/or complex formation with other molecules can stabilize a protein through a reduction of the Gibbs free energy of the complex, resulting from the creation of new molecular interactions (hydrogen bonds, van der Waals interactions, etc.) or conformational reordering of the target protein. This increase in the Gibbs free energy results in an increase in thermal stability and thereby an increase in the melting temperature (Tm). Measurements of the Tm of a protein in the presence and absence of environment changes or ligands result in an estimate of the thermal shift (ΔTm) deriving from these differences (Scott et al. 2016) (Fig. 1b). This shift is typically an indicator of complex formation and/or thermal stabilization. However, it should be borne in mind that while the resulting temperature shift is directly related to the change in the Gibbs free energy, it is a measurement deriving from both binding interactions and any resulting conformational changes in the target protein, and as the thermal stability profile is generated over a temperature range, it is difficult to generate a reliable room temperature dissociation constant (kd = exp −∆G/kT; k = Boltzmann’s constant and T = thermodynamic temperature) directly from ΔTm. However, solely concentrating on Tm may mean that other systemic and thermodynamic information about protein stability can be lost. The propensity of the protein to aggregate in certain conditions is one such factor. An environmental change could result in a difference in aggregation behavior but leaves the Tm unchanged. For an in-depth review on this topic, please see Wakayama et al. (2019).Fig. 1a Typical thermal denaturation profile of a protein sample. Fluorescence emission changes with the temperature. The sigmoidal curve indicates the cooperative unfolding status of the protein from trace amounts of SYPRO Orange (yellow) bound to the native protein (green). The peak indicates that all proteins are unfolded to linear peptides or that the hydrophobic core is exposed to SYPRO Orange. Multiple mechanisms exist for the reduction in fluorescence after the peak, including temperature-driven decrease in the binding constant of the dye (so less dye is bound to the protein), the pocket binding the dye being more mobile (allowing for more quenching by solvent); the dye itself is more mobile such that the degree of planarity required for electron conjugation/aromatic character is lessened and protein aggregation and dye dissociation through the exclusion of the dye from hydrophobic cores. The midpoint of the transition curve is the melting temperature (Tm). b DSF curve showing the unfolding status of a target protein in the absence (blue) and presence (orange) of a ligand. The difference in the melting temperature indicated as ΔTm. c Sample with high background fluorescence at the beginning at lower temperature (red) compared with a typical well-folded sample (blue) in the DSF assay. Improperly folded, aggregated, denatured protein or hydrophobic area such as a lipid bilayer exposed to the dye will cause high background at low temperatures. d Multiple transitions appearing during the heating process can be caused by different domains, aggregation increasing with temperature, or ligands that stabilize a portion of the protein sample (orange); typically one Tm similar to the native protein is accompanied by one or more Tm at a higher temperature during the denaturation. e–g Overview of NanoDSF. e Intrinsic fluorescence of tryptophan is measured at both 330- and 350-nm wavelengths and plotted versus temperature from 20 to 60 °C during unfolding. f F330/350 fluorescence ratio intensity of tryptophan plotted against temperature. g The melting temperature is calculated by the first derivative of the F330/350 plots, with the sample given here showing a Tm of 48 °C. All the figures above represent thermal unfolding curves of the menin protein and are obtained from DSF experiments conducted in our lab. The experiments were performed by using either the Bio-Rad CFX96 Real-Time PCR system or the NanoTemper Prometheus NT.48 system. Curves were plotted from the fluorescence data using ExcelFull size imageIn order to monitor the thermal unfolding transition of target protein in a suitably sensitive but precise way, fluorescence has been used as the response signal. There are two main sources of this fluorescence in use today that may be broadly classed as (i) extrinsic fluorescence and (ii) intrinsic fluorescence.Extrinsic fluorescenceThe fluorescence of extrinsic fluorescent dyes is sensitive to their environment. Typically such dyes are quenched in aqueous solutions with proteins in their native folded state and provide a fluorescence signal only when the target protein begins to unfold. This unfolding allows the freely diffusing dye to interact with the exposed residues of the hydrophobic core (Fig. 1a). This approach relies on the following assumptions (in rough order of frequency as experienced by the authors):

a.

The target proteins do not possess significant hydrophobic patches on their exposed surfaces, the presence of which would lead to increased background in fluorescence (Fig. 1c).

b.

The protein is in a stable state at the beginning of the experiment, and DSF experiments using extrinsic dyes are typically performed at concentrations of 0.1–0.5 mg/ml (0.01–0.1 μM). Aggregation and/or sample instability may lead to the presence of multiple species of target protein within the experiment, both leading to increased fluorescence background from any conformational variability and resulting in variable thermal stability profiles of the different order oligomers (Fig. 1c).

c.

The target protein shows no significant binding interaction(s) with the dye in use—resulting in the shielding of the dye from the aqueous environment prior to protein unfolding and a resulting increase in fluorescence background.

d.

The target protein is composed of a single domain, as the unfolding of distinct domains is likely to occur with different Tm values resulting in a complex thermal stability profile (Fig. 1d). However, while the profile might be more complex, it is often easier to differentiate between the signals from multiple domains and this can provide valuable information as seeing a Tm shift more strongly in a specific domain can provide information about a potential binding site.

e.

No major structural rearrangements of the target protein are provoked by an increased temperature prior to its unfolding, although in such cases, deconvolution of the thermal stability profile may still be possible.

f.

The sample and dye do not chemically react with other components present in the experiment over the temperature range used.

Dyes in common usageThere are many commercial dyes available (Hawe et al. 2008). Dyes such as bis-ANS and Nile Red have been used for decades; the extrinsic dyes are summarized in Table 1. However, these dyes all possess a significant background in the presence of folded proteins. To date, the most favored dye for DSF is SYPRO Orange, mainly owing to its high signal-to-noise ratio (Niesen et al. 2007), as well as its relatively long excitation wavelength (near 500 nm). This minimizes the interference of most small molecules as these typically have absorption maxima at shorter wavelengths.Table 1 Overview of extrinsic fluorescence dyes applied in protein characterizationFull size tableIntrinsic fluorescenceAnother source of fluorescence is from the protein sample itself. In 2010, Schaeffer’s team reported a new method, using green fluorescent protein (GFP) to quantify the stability of a target protein (Moreau et al. 2010). In these experiments, a GFP tag was fused to a protein of interest through a peptide linker and used as a reporter system for protein unfolding and aggregation. The fluorescence signal from the GFP changes based on its proximal environment, meaning its signal can be used to monitor the unfolding of the protein it was linked to. Since GFP only starts losing fluorescence around 75 °C, this approach suits a large number of proteins which are significantly less stable than GFP (Moreau et al. 2010). While this is potentially an elegant solution to remove reliance on a fluorescent dye reporter, there do remain a number of limitations:

a.

The potential for interaction between GFP and the target of interest influencing the target protein conformation, thereby introducing a bias into the measured interactions with ligands.

b.

The potential for a GFP-linked domain to influence the oligomeric state of the target protein—either promoting or inhibiting assembly—with a similar effect on the target protein conformation.

c.

This approach is unsuitable for protein targets that have a similar Tm to that of GFP—in which case the unfolding signal of the target protein will be masked by that of GFP

d.

Ligands that may result in a significant elevation of the target-to-ligand complex Tm will not be clearly observed due to a similar masking effect.

e.

This approach is unable to directly distinguish between compounds that interact with GFP and those that interact with the target protein, although this can be addressed through the use of a GFP only control.

In 2014, a label-free DSF technique marketed as nanoDSF was developed (Alexander et al. 2014). This approach removes the requirement for an extrinsic dye or fusion tag, instead of relying on the change of intrinsic tryptophan fluorescence at 330 nm and 350 nm (Fig. 1e). Unfolding/denaturation results in a change in the microenvironment polarity around tryptophan residues, leading to a redshift of fluorescence (Ghisaidoobe and Chung 2014). In this approach, the Tm can be determined by measuring the ratio of the fluorescence at 330 nm and 350 nm against temperature (Fig. 1f, g). The commercial instrument Prometheus NT.48 (NanoTemper Technologies, Munich) allows a rapid analysis for both ligand screening and buffer composition optimization and, unlike the previous approaches, allows for measurements to be made in detergent-containing solutions—a prerequisite for DSF application to membrane proteins. Due to the nature of extrinsic dyes, which can bind (and fluoresce) in the presence of lipid bilayers and detergent micelles, conventional DSF cannot handle the detergent selection for membrane protein solubilization. The dye-free nanoDSF avoids this problem by using intrinsic fluorescence. Another benefit to intrinsic fluorescence is the ability to observe the transition both from folded to unfolded states and from unfolded to folded states. This allows for the detection of hysteresis (Andrews et al. 2013). The presence of hysteresis can provide information about protein stability (Mizuno et al. 2010). Due to the presence of dye, this is not possible when using an extrinsic fluorescence approach. However, the intrinsic fluorescence method also has several key limitations:

a.

The number of tryptophan residues in the target protein amino acid sequence needs to be considered before adopting this approach, since at least one tryptophan has to be present and the ratio of tryptophan present in the target protein sequence is the limiting factor to detect an unfolding signal.

b.

Experiments that result in complex populations in the thermal profile (e.g., presence of both bound and unbound states—see below) may not be successfully identified due to signal sensitivity.

c.

This approach requires a significantly larger investment for the associated equipment.

Finally, it should be clearly borne in mind that all DSF approaches are sensitive to the intrinsic fluorescence properties of the molecules present in the screen under examination, which can result in a wide variation in the background of thermal profiles—resulting in false negatives. While the use of extrinsic dyes alleviates this to some extent, as the role of the dyes in use is to significantly amplify the unfolding signal, there still remains the potential for screening components to interact with the reporting dye.Recent applications of DSFLigand screening in drug discoveryDetermining the interaction between receptors and members of a small molecule library is addressed by detecting and measuring changes in the physicochemical properties of any ligand-to-target complexes that are formed. Quantitative information arising from receptor-ligand complex formation can then drive the development process through structure-activity relationships (SAR). In the last few years, great efforts have been expended to find a general and universally applicable approach to detect binding (and ideally estimate binding affinity, Kd) between biomolecule receptors and small molecule ligands. As a result, many new biophysical technologies have emerge, briefly:

a.

Differential scanning calorimetry (DSC), which monitors the change in heat capacity of protein samples undergoing temperature-induced melting transitions in the presence and absence of small molecule ligands (Pantoliano et al. 1989).

b.

Isothermal titration calorimetry (ITC), which compares the temperature differences between a reference and receptor solution to quantify the kinetic parameters of binding (Herrera and Winnik 2016).

c.

Surface plasmon resonance (SPR), which records the angular shift of polarized light reflected from a metal film, containing a surface-immobilized target leading to changes in refractive indices upon association and dissociation of ligand (Navratilova and Hopkins 2010).

d.

Microscale thermophoresis (MST), which detects the thermophoretic behaviors of receptors in the presence of ligands under heating in capillaries (Wienken et al. 2010).

e.

NMR-based chemical shift screening, ligand-based or protein-based NMR monitors chemical shift perturbation induced by ligands; thereby, both Kd and the structural conformation of complexes can be determined.

f.

X-ray crystallography–driven fragment optimization based on the electron density of ligands, providing interaction details at atomic resolution.

g.

Mass spectrometry–based approaches, protein samples, and bound ligands are ionized preserving non-covalent interactions. Subsequently, the mass of protein and ligands can be acquired with high accuracy (multiple instances are provided in the table below).

h.

Biolayer interferometry (Wartchow et al. 2011) provides similar binding information to that obtained by SPR, with advantages in signal stability arising from the use of interferometry patterns.

MethodPrincipleAdvantagesLimitationsRefLigand-observed NMRShift change in magnetic state of ligand due to bindingMany fragments can be tested simultaneouslyUses a lot of protein. Limited to fragments with fast exchange with targetKrimm (2017)Protein-observed NMRProtein NMR peak shift induced by bindingAble to determine binding site. Titration possible to determine KDRequires large amounts of protein. Limited throughputKrimm (2017)X-ray crystallographyX-ray diffraction of cocrystallized protein-ligand complex or soaked apo-crystalProvides structural information of ligand-binding mode and interactions with the target. Enables use of computational methods of hit optimizationNeeds good-quality crystals. Not all the ligands can acquire cocrystal structures with protein target. Needs synchrotrons to obtain x-ray diffraction data. Requires large amounts of ligandBadger (2012); Patel et al. (2014)SPRRefractive index change due to ligand binding to immobilized target on sensorAble to easily obtain KD and other kinetic data. Uses very little proteinProtein needs to be able to be immobilizedNeumann et al. (2007); Chavanieu and Pugnière (2016); Huber et al. (2017)DSFThermal stability of protein is increased due to fragment bindingHigh throughput, cheap materials, equipment easy to use and widely availableMany false positives and negatives. Typically only provides a yes/no answer. Requires a dye or intrinsic fluorescenceLo et al. (2004); Douse et al. (2015); Bai et al. (2019)Isothermal titration calorimetry (ITC)Heat of the system changes upon binding eventThermodynamic and binding properties of protein—fragment interaction can directly be obtained. Label-freeUses large amount of protein; low throughputChaires (2008); Ladbury et al. (2010); Renaud et al. (2016)Differential scanning calorimetry (DSC)Amount of heat required to increase temperature of sample changes upon bindingHighly sensitive method. Label-freeUses a lot of protein. Low throughputCooper (2003); Bruylants et al. (2005); Erlanson et al. (2016)Native mass spectroscopy (MS)Mass detection of protein-ligand complex in gas phaseHighly sensitive method. Uses very little protein. Label-free. Provides large amount of information, binding affinity, stoichiometryProtein has to be stable in ESI bufferQin et al. (2015); Pedro and Quinn (2016); Ren et al. (2019)Size exclusion chromatography (SEC) MSIncubation of protein in fragment mixture then separation of bound from unbound molecules by SEC, followed by MS detectionVery high throughput. Easy to perform technique requiring simple LC-MSPotential for false negatives for low affinity binders; these can easily get lost during the SEC stepQin et al. (2015); Chan et al. (2017); Ren et al. (2019)Weak affinity chromatography (WAC) MSSeparation of molecules by affinity to immobilized receptor on the WAC column followed by MS detectionEasy method to use. High throughput possible by using fragment mixturesProtein needs to be immobilized on the column(Duong-Thi et al. 2011; Chan et al. 2017; Ohlson and Duong-Thi 2018)Hydrogen-deuterium exchange (HDX) MSLigand binding affects deuteration rate of protein residues. Which is detectable by massBinding site can directly be elucidated and gives information about protein conformational changesLow throughput and expensiveChan et al. (2017); Marciano et al. (2014)Microscale thermophoresis (MST)Change in the molecular motion of the target in a temperature gradient due to ligand bindingMeasurements can be performed in native buffers. Allows for KD determinationTarget needs to be labeled or have sufficient intrinsic fluorescence. Relatively low throughputLinke et al. (2016); Rainard et al. (2018)Affinity capillary electrophoresis (ACE)Change in electrophoretic mobility of the ligand due to binding to target (in solution)High throughput. Sensitive method. Uses small amounts of protein and ligand. Both target and ligand are free in solutionRequires detectable probe molecule or detectable fragmentsXu et al. (2016); Austin et al. (2012); Farcaş et al. (2017)Biolayer interferometry (BLI)Interference pattern change due to ligand binding to immobilized target on biolayerCan obtain KD and other kinetic parameters. Uses a small amount of proteinImmobilization of protein is requiredWartchow et al. (2011)With the advent of modern advances in bioinformatics and proteomics, many new disease targets have been identified (Lippolis and Angelis 2016). In parallel chemical synthesis, methods are more advanced and refined, being capable of rapidly producing large libraries of diverse compounds. A particularly important subgroup of these methods is those that are compatible with multicomponent reaction (MCR) chemistry (e.g., the UGI reaction) which can generate large libraries of highly specific compounds in a short amount of time. However, the pace at which chemical libraries could be screened using conventional techniques such as NMR and ITC often could not keep up with the speed that the libraries were being created, or the numbers of discrete molecules contained in these libraries.Modern DSF is well placed to address these large and diverse libraries, as it utilizes a real-time PCR machine to rapidly screen multiple molecules at once against the target protein, meaning it can handle the high throughput of compounds much better than many other technologies. With relatively low consumption of protein sample, 96, 384, or 1536 ligands can be analyzed in a single screen that takes ~ 1 h and provides qualitative binding information; it is well-suited for high-throughput library screening. This efficient workflow makes it possible to judge and rank potential binding affinity.In 2001, Pantoliano introduced a DSF-based high-throughput methodology for a variety of therapeutic target proteins (human α-estrogen receptor (ESR), bacteriorhodopsin, human α-thrombin, bovine liver dihydrofolate (DHFR), the extracellular domains of the fibroblast growth factor receptor-1 (D(II)-D(III)FGFR), and the enzyme PilD; Pantoliano et al. 2001). These targets were screened against various small molecules from combinatorial libraries, including known binding ligands. Experiments showed that the Kd calculated from Eq. (1) based on the Tm values obtained experimentally gives very similar values to those previously acquired by other techniques. For example, tamoxifen inhibits the ESR antagonist with an IC50 value reported as 0.42 μM (Bolger et al. 1998), whereas the miniaturized thermal shift assay provided an affinity of 1.1 μM. The known ligand pentosane polysulfate is reported to have a Kd of 11 μM with FGFR-1, as measured by ITC titration (Pantoliano et al. 1994), while the thermal shift assay, i.e., DSF, shows a similar binding ability of 5.5 μM. Thus, the reported thermal shift assay supports a reliable alternative for determining the interactions between proteins and small molecules.$$ {K}_L^{Tm}=\frac{\mathit{\exp}\left\{-\Delta {H}_u^{T0}/R\left[1/{T}_m-1/{T}_0\right]|+\Delta {C}_{pu}^{T0}/R\left[ In\left({T}_m/{T}_0+{T}_0/{T}_m-1\right)\right]\right\}}{\left[{L}_{Tm}\right]}\kern0.5em $$

(1)

where\( {K}_L^{Tm} \):

ligand association constant at Tm

Tm:

midpoint for the protein unfolding transition in the presence of ligand

T0:

midpoint for the unfolding transition in the absence of ligand

\( \Delta {H}_u^{T0} \):

enthalpy of protein unfolding in the absence of ligand at T0

\( \Delta {C}_{pu}^{T0} \):

change in heat capacity on protein unfolding in the absence of ligand

[LTm]:

free ligand concentration at Tm ([LTm]≅ [L]total when [L]total >>[Protein]total)

R:

universal gas constant

DSF has a direct application in fragment-based ligand design (FBLD) due to the ease of use in high-throughput screening. In this approach, small molecule building blocks (100–150 Da) are potentially pooled (3–5 molecules per pool) and screened (Elkin et al. 2015; Valenti et al. 2019). Although these small molecular mass compounds are unlikely to possess high affinity by themselves, this pooled approach allows for a significant reduction in the number of experiments that need to be performed to screen a large library. Successful “hit” pools identified on the basis of a shift in Tm can then be examined in more detail to uniquely identify fragments of interest and hits can be grouped to provide a primary metric for lead compound optimization. This strategy also provides a high probability of adding blocks to the final scaffold of lead compounds (Mashalidis et al. 2013), and two recent examples of the use of DSF in lead discovery are provided below.DSF as a simple and robust mechanism to probe fragment-binding modes and suggests linking strategiesTuberculosis (TB), caused by Mycobacterium tuberculosis (Mtb), remains one of the top 10 causes of death, and Mtb is the leading infectious agent (above HIV/AIDS) worldwide. In 2017, 10 million people developed TB resulting in 1.6 million deaths (World Health Organization 2018). Drug-resistant TB continues to be a public health crisis, and we still lack robust therapies to combat this burden. Consequently, new antitubercular agents that target TB with novel mechanisms are urgently needed. Biotin, also known as vitamin B7, is an essential cofactor for Mtb (Hayakawa and Oizumi, 1987). As Mtb produces biotin in order to support growth and proliferation, but this vitamin is present at very low concentration in human blood (Sassetti and Rubin 2003), therefore, targeting the biotin biosynthesis route intermediate by PLP-dependent transaminase (BioA) turns out to be a promising strategy (Mann and Ploux 2006). Dai and colleagues screened a Maybridge Ro3 fragment library with approximately 1000 compounds against BioA using DSF and discovered 21 “hit” compounds—identified as those that increased the Tm more than 2° (Dai et al. 2015). Subsequent X-ray diffraction data of cocrystals confirmed 6 fragment hits binding within the active site. The binding affinity and ligand efficiencies were cross-validated by ITC, giving a range between 7 and 42 μM in affinity and between 0.43 and 0.55 in ligand efficiencies, respectively. Comparison of all the available hits provided the basis for understanding the interaction mode of residues involved in the active pocket, leaving sufficient guidance for a lead sketch optimization consistent with the active site conformational states. Moreover, the scaffold of the small fragments found by DSF and crystallography also matched existing potent inhibitors previously reported (Park et al. 2015), further demonstrating that this strategy can be a reliable method for ligand screening.The same strategy was implemented by Hung’s team, targeting pantothenate synthetase (PS) of TB (Hung et al. 2009). Pantothenic acid (vitamin B5) plays an important role in fatty acid metabolism. It is formed through condensation of pantoate with β-alanine by pantothenate synthetase (PS), and blocking this pathway will likely impact the growth of Mtb (Sambandamurthy et al. 2002). In fragment screening via DSF, ligand 2 was identified from 1300 fragments with a ΔTm of 1.6 °C (Fig. 2). This was further confirmed by WaterLOGSY NMR spectroscopy and ITC (Kd = 1 mM). The associated X-ray structure showed that 2 binds across the pantoate-binding pocket P1, extending further along the surface of PS, to a point 3.1 Å away from another binding site of ligand 1 in the same pocket. A test with both ligands soaked into crystals showed the presence of both fragments in the active site without clashes, in conformations similar to their individual binding modes (Fig. 2). Therefore, fragment linking and optimization were recruited to enhance binding properties, with different linkers based on the adjacent structures inside the pocket. Subsequently, lead compound 3, which links fragments 1 and 2 by an acyl sulfonamide, showed a 500-fold stronger binding affinity than the individual fragments (Fig. 2).Fig. 2Fragments 1 and 2 soaked as a cocktail into the crystal of pantothenate synthetase. The two fragments are found to bind in distinct positions. Overlay of the linked lead compound 3 with fragments 1 and 2 in the active site of P1 of pantothenate synthetase. Fragments 1 and 2 shown as sticks in green. The benzofuran group is slightly rotated relative to fragment 2, indicating that the stereochemical constraints of the linker do not allow this moiety to adopt its optimum conformation. Figures created by using PyMol, based on PDB entry 3IMG and 3IVX (Hung et al. 2009)Full size imageDSF combined with limited proteolysis in the identification of tankyrase inhibitorsA fragment-based study performed by Larsson in 2013 gives a clear example of how DSF can be used to identify high-quality fragments followed by guiding the construction of a lead compound (Larsson et al. 2013). In this assay, the poly-ADP-ribosylating enzyme tankyrase was screened against a 500-compound fragment library (each present at 1 mM). To avoid oddly behaving compounds and minimize false-positive rates (i.e., pan-assay interfering compounds, PAINS) (Baell and Nissink 2018), identified hits are further validated to genuine “hits” by checking for a dose-dependent DSF response over a range of concentration (from 5 to 4000 μM). In the DSF screening of tankyrase 2, a “hit” melting profile was interpreted as those showing a two/multiple-state transition, which significantly complicated the fitting of Tm for weakly binding fragments (Fig. 3a). After adding chymotrypsin to perform an in situ digestion and remove less-ordered contaminants, they succeeded in simplifying the sigmoidal melting cure (Fig. 3b). Dose-response experiments then validated initial “hits” through an apparent increase in Tm upon elevated concentrations of an initial “hit” (Fig. 3c). Based on the cocrystal structure of TNKS2 with validated hits, various modifications of the hit fragment were proposed and evaluated. The 4-position methyl group was maintained as it protrudes down toward the catalytic glutamate, whereas changes in the 7-position group, which points toward the extended pocket responsible for adenosine binding, showed distinct differences when ligated to different functional groups. Starting with an initial fragment of 12-μM affinity, multiple rounds of modification and validation by DSF, SPR, enzymatic activity (IC50), and X-ray crystallography yielded a lead compound with an inhibition activity (IC50) of 9 nM and binding affinity (Kd) of 16 nM against TNKS2. The elegant approach of limited proteolysis of the less stable (i.e., unbound) form of the target directly addresses one limitation of DSF—incomplete binding leading to multiple transitions in the thermal profile—amplifying weak binding. However, it is likely that such an approach will be highly dependent on the target under examination and may not be generally applicable.Fig. 3a Tankyrase 2 melting curves without chymotrypsination in the absence (black) and presence (red) of a stabilizing fragment. b Tankyrase 2 melting curves treated with chymotrypsin in the absence (black) and presence (red) of the same stabilizing fragment. c Concentration-dependent response for the stabilizing fragment with chymotrypsin-digested tankyrase. d The workflow of the final lead compound optimization from the initial hit to the end was guided by DSF. This figure was adapted with permission from Larsson et al. (2013). Copyright 2013 American Chemical SocietyFull size imageIn summary, the examples above both show that fragment-based drug discovery (FBDD) has become a mainstream choice for high-throughput screening for lead discovery of therapeutic interest (Congreve et al. 2008; Murray and Rees 2009) and that DSF has been validated as a robust option in preliminary screening in FBDD for more than 2 decades (Pantoliano et al. 1997). The use of DSF in fragment screening is facilitated by its low sample consumption—both in proteins and in chemicals—as well as the rapid determination of experimental ΔTm determination—reducing labor-intensive work and providing simplified screening protocols.The use of DSF in buffer screening and optimization of protein stability and crystallizationIn proteomics studies, inter-related biochemical, cellular, and physiological information is essential to reveal protein mechanisms. A major source of information is the use of structural, functional, and chemical genomics to characterize target proteins (Christendat et al. 2000). However, the common first step for all these approaches is the purification of the target protein, which remains challenging in many cases. On average, only 50–70% soluble protein and 30% membrane proteins from prokaryotes can be expressed in a recombinant form, and among those successfully expressed, only 30–50% can be purified in a homogeneous state (Christendat et al. 2000; Norin and Sundström 2002; Dobrovetsky et al. 2005). Eukaryotic proteins—including many biomedically interesting targets from humans—seem even more challenging (Banci et al. 2006).Traditional solutions for protein production and purification mainly rely on the screening of recombinant hosts, encoding construct sequences, expression conditions, and then purification conditions (Gräslund et al. 2008; Rosano and Ceccarelli 2014; Wingfield 2015). In the last two steps, the addition of specific additives or changing buffer composition can significantly increase the solubility of recombinant proteins, as well as improving the thermal stability of the target to prevent protein unfolding or aggregation—even at a low temperature. There have been many reports (Sarciaux et al. 1999; Vedadi et al. 2006; Reinhard et al. 2013) showing that optimization of the purification conditions results in enhanced protein stability or solubility and it is not unreasonable to propose that buffer optimization should be seen as an integral part of any research project that relies on isolated protein samples. Even minor gains in protein stability can be significant in the context of process engineering, for example in the mass production of antibodies for therapeutic purposes.One remarkable case is that of the recombinant protein dnaB, produced in E. coli. Initially, it was shown to be highly unstable in the purification buffer—even when stored at 0 °C, 90% enzymatic activity was lost within 30 min. In a stepwise screening process where specific chemical reagents (Mg2+, ADP, (NH4)2SO4, and glycerol) were added, 90% activity was retained after extensive storage at 60 °C in the optimal buffer. Furthermore, the new buffer helped the isolation of soluble dnaB at increased yields and subsequent crystallization (Arai et al. 1981). While this is undoubtedly an extreme example, this clearly shows the value of buffer optimization.In the early years of structural genomics, a generally applied strategy was to use a default purification buffer for the majority of protein targets, with detailed optimization of sample buffer performed only to address pathological issues (aggregation, loss of activity, change in oligomeric state, etc.) (Mezzasalma et al. 2007). As shown below, this likely impacted the ultimate success of structural genomics projects, in which the growth of high-quality crystals from purified samples represented the major bottleneck. To address the issue of buffer optimization, Ericsson and coworkers developed a DSF-based screening system (comprised of different pH buffers, additives, heavy atoms, etc.) to test 25 different proteins expressed in Escherichia coli (Ericsson et al. 2006). The buffers consisted of a set of 23 different buffering agents at a concentration of 100 mM with a pH range from 4.5 to 9.0. Because each pH step is only 0.2 to 0.5 pH unit, it makes the screen wide enough for the majority of proteins investigated currently.In some cases, protein Tm was dramatically influenced by a single pH buffer, correlated with a preference for specific ionic effects. For example, at pH 7, the Tm of protein AC07 in K-phosphate is 37 °C, whereas it is 46 °C in the presence of Na-phosphate (Fig. 4a). In order to decouple the influence of the choice of buffer and the final pH, a three-component buffer system (Newman 2004) was implemented, which allowed a wide range pH without altering the composition of buffer chemicals. The citric acid-Hepes-Ches (CHC) buffer, which covers the pH range from 4 to 10, can quickly identify the most favorable pH of target proteins. This work showed that the Tm of the targets examined followed a typical bell-shaped curve. For example, AD28 demonstrated lower temperature stability values at both low and high pH (pH = 4 and 10), with a maximum stability close to pH 6.4.Fig. 4a Unfolding temperature of AC07 in various pH buffers of different compositions. Na-phosphate (red bar) and K-phosphate (blue bar) at a pH close to 7.4 showed a significant difference in Tm. b Melting temperature curves of the protein AD21 screened against different additives. As an essential chemical needed in the proline biosynthetic pathway, NAD(P)H (yellow) showed a visible increase in thermal stability when incubated with the target protein. The figures are adapted from Ericsson et al. (2006). Copyright 2006 with permission from ElsevierFull size imageCombinations of the above buffer optimization with additives, such as heavy metals, or substrates/cofactors like NADH at optimal pH can further enhance protein thermal stability. For example, the addition of NADH was found to increase the melting temperature of AD21 significantly (ΔTm ≈ 20 °C; Fig. 4b), which correlated with the previously known fact that it is an essential cofactor of AD21 in the catalysis of the last step in proline biosynthesis.In summary, DSF screening of additives provided data to optimize the buffer conditions for crystallization screening (Reinhard et al. 2013). Additives that gave a positive thermal shift (Tm) compared with control samples increased the protein crystallizing rate by 70%, while additives that showed destabilizing effects reduced the chance of getting crystals by around 50% compared with the control buffer. This observation strongly suggests a correlation between protein stability/solubility and crystallogenesis. For excellent in-depth reviews into the use of DSF to optimize crystallization buffers, the reader is referred to Boivin et al. (2013) and Reinhard et al. (2013).Structural biology plays an important role in early-stage drug discovery, as the elucidation of the binding modes of “hit” compounds can provide essential information to drive downstream, lead compound development (de Kloe et al. 2009; Wang et al. 2019). While crystallization of proteins relies on a number of sample properties, with sample purity and homogeneity generally agreed to be the key determining factors (Giegi et al. 1994; Dale et al. 2003; Ericsson et al. 2006), thermal stability has also been shown to be a critical parameter in a successful outcome during crystallogenesis. In a study carried out by Dupeux et al. (2011), 657 different proteins were screened by DSF, then subjected to automated vapor-diffusion crystallization. Based on an analysis of the protein melting point (Tm) and visually determined crystallization hits, the authors were able to draw clear inferences on the importance of thermal stability on the crystallization process. In this study, 437 of the 657 samples unfolded show clear and sharp temperature transitions. This behavior may be interpreted as the result of a sample population consisting of a single overall conformation, with relatively little conformational fluctuation around the “mean” fold—a scenario which is likely to be more conducive to crystallization than a sample with a high degree of conformational variation due to thermal mobility of its component elements. The average Tm for the ensemble of samples was 51.5 °C over a range of 25 to 95 °C (Fig. 5). Notably, proteins with a Tm of 45 °C or higher displayed a greater tendency to crystallize when incubated at 20 °C, with successful crystallization outcomes of 49.1%. For proteins with a Tm below 45 °C, the likelihood of crystal growth chance at 20 °C dropped to 26.8%. Additionally, a number of proteins with a Tm between 25 and 45 °C produced crystals at the lower temperature of 5 °C, where crystallization was initially unsuccessful at 20 °C. The study confirmed a previous observation that thermophilic proteins have higher rates of crystallization than those from mesophilic organisms, despite similar Tm values. In addition, a report from Szilágyi also implied that thermophilic proteins have a lower proportion of unstructured regions (Szilágyi and Závodszky 2000), with the inference that the disordered regions will hamper crystallization.Fig. 5Tm and success rate in crystallization: all the samples were incubated for crystallization at 20 °C; the numbers above the bars indicate the success rate in crystallization of each class. The samples from extremophilic organisms consist of 12 proteins with Tm between 70 and 95 °C. The figure is adapted from Dupeux et al. (2011). Reproduced with permission of the International Union of CrystallographyFull size imageAs the thermal stability of a sample may influence its chances of crystallizing, it becomes clear that optimizing the sample buffer in which the protein is finally purified and concentrated prior to crystallization can provide benefit to structural biologists, and structure-based drug design in particular. In a typical DSF buffer screening experiment, the conditions (buffering agent, pH, additives, etc.) that result in the largest thermal shifts are often combined and the resulting buffer is then used for purification and crystallization. However, this process can be complicated when multiphasic unfolding behavior is encountered as it makes accurate Tm determination more difficult. A multiphasic unfolding curve typically indicates either the presence of multiple, independently folding, domains (Ionescu et al. 2008) or a heterogeneous state of the protein sample in solution (Choudhary et al. 2017), or ligand binding is not fully saturated with protein targets (Shrake and Ross 1992; Matulis et al. 2005), which may disrupt crystallogenesis and hinder protein functional characterization. Here, DSF can also be applied to guide sample preparation buffer screening for crystallization by replacing the buffer ingredients or ligands stepwise. Geders et al. reported a multiphasic unfolding behavior when his team attempted to crystallize pyridoxal 5-phosphate (PLP)–dependent transaminase BioA from Mycobacterium tuberculosis (Geders et al. 2012). During buffer optimization for crystallization, BioA displayed a multiphasic unfolding behavior without PLP; subsaturation of cofactors in the protein-cofactor system also yields a biphasic melting curve. The protein heterogeneity resulting from insufficient levels of cofactor PLP could potentially impact crystallization. To avoid the competition for PLP binding by other factors and to induce PLP saturation of BioA, DSF was used to study PLP binding. The initial buffers used in both lysis and purification (Dey et al. 2010) were Tris-based—generating a tri-phasic melting temperature curve with transitions at 45, 68, and 86 °C (corresponding to misfolded, apo, and PLP-bound BioA, respectively (Fig. 6a)). The sample also displayed significant precipitation at higher concentration levels. The electron density from a crystal grown from a Tris buffer showed no interpretable density for a bound PLP molecule. Replacing the Tris buffer with Hepes within the purification (both lysis buffer and final purification buffer) resulted in a decreased tendency for multiphasic melting curves, especially while Hepes completely replaced Tris in both lysis and purification buffer (Fig. 6b). This result suggested that the Tris buffer partially degraded the PLP, resulting in unsaturated PLP binding to BioA partially. This partial degradation was further supported by a UV-Vis spectroscopy assay, in which PLP in Tris buffer showed an absorbance maximum near 420 nm, similar to that shown by PLP in the Schiff base form instead of a free aldehyde (Fig. 6d). PLP in Hepes buffer showed absorbance at 390 nm, similar to that of PLP in water. By replacing Tris with Hepes in all purification buffers and adding increased concentrations of PLP, the multiphasic melting curves were replaced with a single, sharp transition curve with a Tm at 88 °C. These optimizations also improved the size and quality of the crystals obtained and also resulted in clear electron density for a bound PLP molecule. Thus, the DSF analysis correlated with heterogeneity and suboptimal crystallization outcomes. This example also highlights two complications in small molecule screening: firstly, the use of Tris (or primary amines which can form Schiff base with aldehydes) should be avoided with PLP-dependent proteins—and researchers should be aware of the potential for similar effects in other protein cofactors. Secondly, care should be taken when analyzing multiphasic DSF profiles, as they may be due to molecular interactions of the screen with the buffer, rather than the protein target.Fig. 6a DSF melting curves of BioA with PLP and Tris in both lysis and storage buffer, which shows multiple peaks during denaturing. b A sharp DSF melting curve of BioA with subsaturation of PLP; misfolded and apo peaks were eliminated after BioA was saturated with PLP, resulting in enhanced stability of BioA, with a Tm at 88 °C. c First derivative overlap of the corresponding melting curves. The red line indicates BioA in Tris buffer, with multiple transitions at 45, 68, and 86 °C, representing the misfolded, apo, PLP-bound BioA, respectively. The blue line represents BioA saturated with PLP for which the Tm was enhanced dramatically to 88 °C. d UV-Vis spectroscopy of PLP or PLP-BioA(holo) at various conditions; 400 μM PLP in water (cyan) has the same absorbance as in Hepes buffer (brown); PLP-bound BioA(holo) (purple) showed the same absorbance close to 420 nm as PLP in Tris buffer (black). The figures are adapted from Geders et al. (2012). Reproduced with permission of the International Union of CrystallographyFull size imageIn biochemical or biomedical research, a well-folded protein structure with the correct activity is one of the critical factors for in vitro experiments. While numerous recombinant technologies exist to express proteins, greatly facilitating the understanding of proteomics in both prokaryotic and eukaryotic cells, the lack of suitable chaperones in E. coli (the most commonly used recombinant source) results in ~ 80% of these proteins misfolding into insoluble inclusion body without a defined fold or biological activity (Carrió and Villaverde 2002; Sørensen and Mortensen 2005; Gräslund et al. 2008; Rosano and Ceccarelli 2014). Moreover, refolding of proteins from inclusion bodies is an empirical art, with functionally related proteins of different construct designs or from different sources requiring significantly different conditions to support refolding. Thus, systematic and high-throughput compatible assays are needed to address this. In 2016, Biter and colleagues established a DSF-guided refolding method (DGR) to rapidly screen for the refolding of inclusion bodies, including proteins that contain disulfide bonds and novel structures with no preexisting model (Biter et al. 2016). The refolding trials used a PACT (pH, anion, cation testing) sparse matrix crystallization, leveraging the sparse matrix search of buffers to examine the large chemical space of biologically compatible buffers. Inclusion bodies were purified by centrifugation prior to solubilization in chaotropes (urea or guanidine) and the addition of a fluorescent dye (SYPRO Orange). Precipitants were excluded from the screen (Fig. 7a). The solubilized targets were incubated with components of the PACT screen for 2 h, centrifuged to remove any resultant precipitation/aggregation, and directly analyzed using DSF. Fluorescence data showing protein unfolding under DSF conditions was interpreted as corresponding to a condition that supported protein refolding. Due to the wide range in pH, cations, and anions, the PACT screen provided clear hints for pepsin refolding (Fig. 7c, d). For disulfide-containing proteins, such as lysozyme, the PACT screen conditions were supplemented with oxidized and reduced glutathione. The resulting thermal melting profile of the refolded lysozyme showed a clear Tm at 65 at pH 9 in the presence of equimolar GSH and GSSH.Fig. 7a The modified PACT screen in use in a refolding assay; three main parts consist of pH screen, cations, and anions in different combinations; the color indicates the Tm found in certain conditions. b Thermal melting profiles of pepsin in native, denatured, refolded, and misfolded states. c Peak height Tm in the PACK screen profile; the color indicates that under acid conditions, pepsin has a higher Tm. d First derivatives of pepsin from the guanidine-solubilized dilution; populations in red correspond to the misfolded state, and blue is natively a folded state. The figures are adapted from Biter et al. (2016)Full size imageAttempts to refold the novel proteins from inclusion bodies also succeeded in generating an improved yield of fibroblast growth factors 19 and 21, leading to crystals. When DGR was applied to the hormone Irisin, the success in refolding helped to generate an eight-dimer crystal form (Schumacher et al. 2013).One year later, colleagues in our group expanded the DGR approach by investigating the refolding agent arginine and other additives in systematic buffer screens (Wang et al. 2017). Arginine has been widely used to suppress protein aggregation in refolding, and it can slow or prevent protein association reactions via weak interactions with the targets (Baynes et al. 2005; Arakawa et al. 2007), distinct from chaotropes such as urea or guanidine. Therefore, we designed two sequential screening kits to provide a general screening strategy. The primary screen is a combination of various pH buffers in the presence or absence of arginine at a concentration of 0.4 M. This can rapidly identify a suitable refolding pH while also screening for the effect of arginine in refolding. A secondary screen is then explored, by adding different sugars, detergents, osmolytes, PEGs, amino acids, concentration gradients of salt, and reducing agents, expanding on the PACT screen which mainly focuses on pH, anions, and cations (Fig. 8). This approach identified optimal refolding buffers for four different therapeutic target proteins from inclusion bodies expressed in E. coli, as well as identifying a final gel filtration buffer for storage or crystallization. A number of factors that affect protein refolding were revealed during this study, including the chemical composition of the buffer, refolding time, redox state, and the use of arginine as an inhibitor of aggregation. For example, DGR analysis of the refolding of interleukin-17A (IL-17A) gave obvious melting transition signals at pH 9.5 in CHC and CHES buffer—but not the MMT or MIB buffer system at the same pH—indicating that the compositions of the buffer have a significant effect. In the presence of arginine, the Tm increased from 40 to 60 °C, suggesting a more stable final product of the refolding process (Fig. 9). Refolding time also plays an essential role in all the assays, as data showed for all the proteins tested that the maximum efficiency appeared at a defined refolding time. The receptor-binding domain of hemagglutinin (HA-RBD) showed a clear melting curve when refolding was limited to 1 h, whereas the melting transition signal disappeared after 6-h incubation in refolding buffer. IL-17A needed extensive refolding time, requiring 15 h for an optimal DGR signal. Additionally, this data demonstrated that buffers optimized from the refolding process are not necessarily ideal for subsequent storage or crystallization—potentially as they stabilize an intermediate in the refolding process, rather than the final folded form.Fig. 8The composition of the secondary additive screen covers a wide range of sugars, detergents, salts, buffers, and reducing agents. This figure is adapted from Wang et al. (2017)Full size imageFig. 9Melting transition of IL-17A in CHC buffer system at pH 9–10 in the absence (a) and presence (b) of arginine; both showed a typical sigmoidal melting curve at pH 9.5. The figures are adapted from Wang et al. (2017)Full size imageDSF applications for in vivo ligand: target interaction validationA common issue in monitoring drug binding and efficacy during therapy is that the interactions between target proteins and drugs cannot be measured directly in cells and tissues. Validation methods normally study downstream cellular responses after multiple doses. Furthermore, some drugs tested may have good binding activity when incubated with target proteins but fail in clinical trials, with later research showing them to not act on the predicted target within cells (Auld et al. 2009; Schmidt 2010; Guha 2011). In 2013, Molina et al. (2013) introduced a new way to monitor the drug interactions inside cells by performing thermal shift assays on cells, lysates, or tissues, which is also based on ligand-induced thermal stabilization of target proteins, but no protein purification steps are needed. The cellular thermal shift assay (CETSA) functions by heating cells, whereby the proteins inside also unfold and precipitate—similarly to the in vitro approaches described above. After extract and centrifugation, the remaining soluble proteins were separated from the precipitate and quantified by Western blotting. Plotting the amount of soluble protein based on the Western blot signal strength provides the CETSA melting curve. In the preliminary study, dihydrofolate reductase (DHFR) and thymidylate synthase (TS) were selected as targets for the antifolate cancer drugs methotrexate and raltitrexed. Samples were exposed to either of the two drugs either as intact cells or as lysates. The result showed a distinct thermal shift increase for DHFR- or TS-treated cells compared with controls. To investigate drug concentration effects, an isothermal dose-response (ITDR) method has also been developed to assess binding of compounds. In this approach, cell lysate is aliquoted and exposed to different serial concentrations of the drug, while keeping the temperature and heating time constant. Following Western blotting, the signal strength can indicate when a higher drug concentration is needed for saturation, which is potentially more useful than commonly used half-saturation points (i.e., IC50, Kd) which are related to affinity. Further research validated that the CETSA method can be applied as a reliable biophysical technique for studies of ligand binding to proteins in cells and lysates. In a recent report, Maji’s group screened a library with more than 2000 small molecules in order to identify inhibitors of CRISPR-Cas9, which could then be used for the precise control of CRISPR-Cas9 in genome engineering. CETSA was used to confirm a hit compound that disrupted the SpCas9:DNA interaction and decreased the Tm of SpCas9 by ~ 2.5 °C in compound-treated cells (Maji et al. 2019). In another structure-based design of a small molecule to target the interaction of menin-MLL in leukemia, an irreversible, highly potent chemical M-525 was also confirmed by CETSA in a cellular assay (Xu et al. 2018). The covalent-binding compound enhanced the thermal stability of menin in both MV4;11 and MOLM-13 cells; the concentration of M-525 used here was as low as 0.4–1.2 nM. Furthermore, CETSA also showed that the compound specifically targeted menin, and no effect was detected on another MLL-binding protein WDR5.ConclusionDSF constitutes a robust biophysical technique for studying protein stability in a particular environment, either within selected buffer conditions or when (partially) saturated with ligands of interest. The protein unfolding thermodynamic parameter ΔTm is monitored as the primary indicator to justify stability changes of the target protein, no matter whether targets were in a purified form, in lysate, cells, or even tissues. Newly emerged label-free nanoDSF approaches especially obviate the need for dyes, allowing the same approach to be applied to membrane protein research, simultaneously addressing problems caused by the interaction between dye and the hydrophobic surface of proteins, or the detergent additives applied and interactions between the dye and other molecules in a screen. Over the almost two decades since it first appeared, the DSF technique has been used to characterize the thermal properties of numerous proteins, aided by low sample consumption and high throughput—making DSF suitable for optimizing buffer ingredients in crystallization, as well as screening large ligand libraries. In terms of ligand-binding validation, although many successful cases have been reported in the literature, it is still important to be aware that this correlation typically occurs for similarly structured compounds within a series, and stubbornly pursuing fragment hits on the basis of significant thermal shifts may mislead further optimization. It should also be borne in mind that ligands can interplay with both the folded and unfolded states of target proteins, and a negative shift in melting temperature does not exclude binding to the native state. Unlike titration-based techniques such as ITC, MST, and SPR in which interaction behaviors of receptors rely on different serial concentrations of ligands and end-point measurements, DSF is sensitive to all stages along a binding pathway, complicating its use to determine the affinity of molecules toward mobile protein receptors. Nevertheless, the robustness and applicability of DSF to address various problems across such a wide range of sample types should ensure its status as a central technology of modern drug discovery.

ReferencesAlexander CG, Wanner R, Johnson CM et al (2014) Novel microscale approaches for easy, rapid determination of protein stability in academic and commercial settings. Biochim Biophys Acta - Proteins Proteomics 1844:2241–2250. https://doi.org/10.1016/j.bbapap.2014.09.016

Article 

CAS 

Google Scholar 

Alexandrov AI, Mileni M, Chien EYT et al (2008) Microscale fluorescent thermal stability assay for membrane proteins. Structure 16:351–359. https://doi.org/10.1016/j.str.2008.02.004

Article 

CAS 

PubMed 

Google Scholar 

Andrews BT, Capraro DT, Sulkowska JI, Onuchic JN, Patricia AJ (2013) Hysteresis as a marker for complex, overlapping landscapes in proteins. J Phys Chem Lett 4:180–188. https://doi.org/10.1021/jz301893w

Article 

CAS 

PubMed 

Google Scholar 

Arai K, Yasuda S, Kornberg A (1981) Mechanism of dnaB protein action. I. Crystallization and properties of dnaB protein, an essential replication protein in Escherichia coli. J Biol Chem 256(10):5247–5252Arakawa T, Ejima D, Tsumoto K et al (2007) Suppression of protein interactions by arginine: a proposed mechanism of the arginine effects. Biophys ChemAuld DS, Thorne N, Maguire WF, Inglese J (2009) Mechanism of PTC124 activity in cell-based luciferase assays of nonsense codon suppression. Proc Natl Acad Sci U S A. https://doi.org/10.1073/pnas.0813345106

Article 

CAS 

Google Scholar 

Austin C, Pettit SN, Magnolo SK et al (2012) Fragment screening using capillary electrophoresis (CEfrag) for hit identification of heat shock protein 90 ATPase inhibitors. J Biomol Screen 17:868–876. https://doi.org/10.1177/1087057112445785

Article 

PubMed 

Google Scholar 

Badger J (2012) Crystallographic fragment screening. In: Methods in molecular biology (Clifton, N.J.) 841:161–77. https://doi.org/10.1007/978-1-61779-520-6_7

Google Scholar 

Baell JB, Nissink JWM (2018) Seven year itch: pan-assay interference compounds (PAINS) in 2017 - utility and limitations. ACS Chem Biol 13:36–44. https://doi.org/10.1021/acschembio.7b00903

Article 

CAS 

PubMed 

Google Scholar 

Bai N, Roder H, Dickson A, Karanicolas J (2019) Isothermal analysis of ThermoFluor data can readily provide quantitative binding affinities. Sci Rep 9:2650. https://doi.org/10.1038/s41598-018-37072-x

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Banci L, Bertini I, Cusack S et al (2006) First steps towards effective methods in exploiting high-throughput technologies for the determination of human protein structures of high biomedical value. Acta Crystallogr Sect D Biol Crystallogr 62:1208–1217. https://doi.org/10.1107/S0907444906029350

Article 

CAS 

Google Scholar 

Baynes BM, Wang DIC, Trout BL (2005) Role of arginine in the stabilization of proteins against aggregation. Biochemistry 44:4919–4925. https://doi.org/10.1021/bi047528r

Article 

CAS 

PubMed 

Google Scholar 

Biter AB, De La Peña AH, Thapar R et al (2016) DSF guided refolding as a novel method of protein production. Sci Rep 6:1–9. https://doi.org/10.1038/srep18906

Article 

CAS 

Google Scholar 

Boivin S, Kozak S, Meijers R (2013) Optimization of protein purification and characterization using Thermofluor screens. Protein Expr Purif 91:192–206. https://doi.org/10.1016/j.pep.2013.08.002

Article 

CAS 

PubMed 

Google Scholar 

Bolger R, Wiese TE, Ervin K et al (1998) Rapid screening of environmental chemicals for estrogen receptor binding capacity. Environ Health Perspect. https://doi.org/10.1289/ehp.98106551

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Bouvier M, Wiley DC (1994) Importance of peptide amino and carboxyl termini to the stability of MHC class I molecules. Science 80. https://doi.org/10.1126/science.8023162

Article 

CAS 

PubMed 

Google Scholar 

Bowling JJ, Shadrick WR, Griffith EC, Lee RE (2016) Going small: using biophysical screening to implement fragment based drug discovery. In: Chen T, Chai SC (eds) Special topics in drug discovery. IntechOpen. https://doi.org/10.5772/66423

Google Scholar 

Bruylants G, Wouters J, Michaux C (2005) Differential scanning calorimetry in life science: thermodynamics, stability, molecular recognition and application in drug design. Curr Med Chem 12:2011–2020Article 

CAS 

PubMed 

Google Scholar 

Carrió MM, Villaverde A (2002) Construction and deconstruction of bacterial inclusion bodies. J Biotechnol 96:3–12. https://doi.org/10.1016/S0168-1656(02)00032-9

Article 

PubMed 

Google Scholar 

Chaires JB (2008) Calorimetry and thermodynamics in drug design. Annu Rev Biophys 37:135–151. https://doi.org/10.1146/annurev.biophys.36.040306.132812

Article 

CAS 

PubMed 

Google Scholar 

Chan DS-H, Whitehouse AJ, Coyne AG, Abell C (2017) Mass spectrometry for fragment screening. Essays Biochem 61:465–473. https://doi.org/10.1042/EBC20170071

Article 

PubMed 

Google Scholar 

Chavanieu A, Pugnière M (2016) Developments in SPR fragment screening. Expert Opin Drug Discov 11:489–499. https://doi.org/10.1517/17460441.2016.1160888

Article 

CAS 

PubMed 

Google Scholar 

Choudhary D, Kumar A, Magliery TJ, Sotomayor M (2017) Using thermal scanning assays to test protein-protein interactions of inner-ear cadherins. PLoS One 12:1–20. https://doi.org/10.1371/journal.pone.0189546

Article 

CAS 

Google Scholar 

Christendat D, Yee A, Dharamsi A et al (2000) Structural proteomics of an archaeon. Nat Struct Biol 7:903–909. https://doi.org/10.1038/82823

Article 

CAS 

PubMed 

Google Scholar 

Congreve M, Chessari G, Tisi D, Woodhead AJ (2008) Recent developments in fragment-based drug discovery. J Med ChemCooper MA (2003) Label-free screening of bio-molecular interactions. Anal Bioanal Chem 377:834–842. https://doi.org/10.1007/s00216-003-2111-y

Article 

CAS 

PubMed 

Google Scholar 

Dai R, Geders TW, Liu F et al (2015) Fragment-based exploration of binding site flexibility in Mycobacterium tuberculosis BioA. J Med Chem 58:5208–5217. https://doi.org/10.1021/acs.jmedchem.5b00092

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Dale GE, Oefner C, D’Arcy A (2003) The protein as a variable in protein crystallization. J Struct Biol 142:88–97. https://doi.org/10.1016/S1047-8477(03)00041-8

Article 

CAS 

PubMed 

Google Scholar 

Dey S, Lane JM, Lee RE, Rubin EJ, Sacchettini JC (2010) Structural characterization of the mycobacterium tuberculosis biotin biosynthesis enzymes 7,8-diaminopelargonic acid synthase and dethiobiotin synthetase. Biochemistry 49:6746–6760. https://doi.org/10.1021/bi902097j

Article 

CAS 

PubMed 

Google Scholar 

Dobrovetsky E, Ming LL, Andorn-Broza R et al (2005) High-throughput production of prokaryotic membrane proteins. J Struct Funct Genom. https://doi.org/10.1007/s10969-005-1363-5

Article 

CAS 

PubMed 

Google Scholar 

Douse CH, Vrielink N, Wenlin Z et al (2015) Targeting a dynamic protein-protein interaction: fragment screening against the malaria myosin A motor complex. ChemMedChem 10:134–143. https://doi.org/10.1002/cmdc.201402357

Article 

CAS 

PubMed 

Google Scholar 

Duong-Thi M-D, Meiby E, Bergström M et al (2011) Weak affinity chromatography as a new approach for fragment screening in drug discovery. Anal Biochem 414:138–146. https://doi.org/10.1016/j.ab.2011.02.022

Article 

CAS 

PubMed 

Google Scholar 

Dupeux F, Röwer M, Seroul G et al (2011) A thermal stability assay can help to estimate the crystallization likelihood of biological samples. Acta Crystallogr Sect D Biol Crystallogr 67:915–919. https://doi.org/10.1107/S0907444911036225

Article 

CAS 

Google Scholar 

Elkin LL, Harden DG, Saldanha S et al (2015) Just-in-time compound pooling increases primary screening capacity without compromising screening quality. J Biomol Screen 20:577–587. https://doi.org/10.1177/1087057115572988

Article 

CAS 

PubMed 

Google Scholar 

Ericsson UB, Hallberg BM, DeTitta GT et al (2006) Thermofluor-based high-throughput stability optimization of proteins for structural studies. Anal Biochem. https://doi.org/10.1016/j.ab.2006.07.027

Article 

CAS 

PubMed 

Google Scholar 

Erlanson DA, Fesik SW, Hubbard RE et al (2016) Twenty years on: the impact of fragments on drug discovery. Nat Rev Drug Discov 15:605–619. https://doi.org/10.1038/nrd.2016.109

Article 

CAS 

PubMed 

Google Scholar 

Farcaş E, Bouckaert C, Servais A-C et al (2017) Partial filling affinity capillary electrophoresis as a useful tool for fragment-based drug discovery: a proof of concept on thrombin. Anal Chim Acta 984:211–222. https://doi.org/10.1016/j.aca.2017.06.035

Article 

CAS 

PubMed 

Google Scholar 

Geders TW, Gustafson K, Finzel BC (2012) Use of differential scanning fluorimetry to optimize the purification and crystallization of PLP-dependent enzymes. Acta Crystallogr Sect F Struct Biol Cryst Commun 68:596–600. https://doi.org/10.1107/S1744309112012912

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Ghisaidoobe ABT, Chung SJ (2014) Intrinsic tryptophan fluorescence in the detection and analysis of proteins: a focus on Förster resonance energy transfer techniques. Int J Mol Sci 15:22518–22538. https://doi.org/10.3390/ijms151222518

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Giegi BYR, Lorber B, Obald-dietrich ATHI (1994) Fifth international conference on crystallization of biological macromolecules, San Diego, California, USA, 8-13 August, 1993. Acta Crystallogr Sect D Biol Crystallogr D50:339–366. https://doi.org/10.1107/S0907444994001344

Article 

Google Scholar 

Gräslund S, Nordlund P, Weigelt J et al (2008) Protein production and purification. Nat Methods 5:135–146. https://doi.org/10.1038/nmeth.f.202

Article 

PubMed 

Google Scholar 

Greenspan P, Mayer EP, Fowler SD (1985) Nile red: a selective fluorescent stain for intracellular lipid droplets. J Cell Biol 100:965–973. https://doi.org/10.1083/jcb.100.3.965

Article 

CAS 

PubMed 

Google Scholar 

Grillo AO, Edwards KLT, Kashi RS et al (2001) Conformational origin of the aggregation of recombinant human factor VIII. Biochemistry 40:586–595. https://doi.org/10.1021/bi001547t

Article 

CAS 

PubMed 

Google Scholar 

Guha M (2011) PARP inhibitors stumble in breast cancer. Nat Biotechnol 29:373–374. https://doi.org/10.1038/nbt0511-373

Article 

CAS 

PubMed 

Google Scholar 

Hawe A, Sutter M, Jiskoot W (2008) Extrinsic fluorescent dyes as tools for protein characterization. Pharm Res 25:1487–1499. https://doi.org/10.1007/s11095-007-9516-9

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Hayakawa K, Oizumi J (1987) Determination of free biotin in plasma by liquid chromatography with fluorimetric detection. J Chromatogr 413:247–250Article 

CAS 

PubMed 

Google Scholar 

Herrera I, Winnik MA (2016) Differential binding models for direct and reverse isothermal titration calorimetry. J Phys Chem B. https://doi.org/10.1021/acs.jpcb.5b09202

Article 

CAS 

PubMed 

Google Scholar 

Huber S, Casagrande F, Hug MN et al (2017) SPR-based fragment screening with neurotensin receptor 1 generates novel small molecule ligands. PLoS One 12:e0175842. https://doi.org/10.1371/journal.pone.0175842

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Hung AW, Silvestre HL, Wen S et al (2009) Application of fragment growing and fragment linking to the discovery of inhibitors of mycobacterium tuberculosis pantothenate synthetase. Angew Chemie - Int Ed. https://doi.org/10.1002/anie.200903821

Article 

CAS 

PubMed 

Google Scholar 

Huynh K, Partch CL (2015) Analysis of protein stability and ligand interactions by thermal shift assay. Curr Protoc protein Sci. https://doi.org/10.1002/0471140864.ps2809s79

Ionescu RM, Vlasak J, Price C, Kirchmeier M (2008) Contribution of variable domains to the stability of humanized IgG1 monoclonal antibodies. J Pharm Sci 97:1414–1426. https://doi.org/10.1002/jps

Article 

CAS 

PubMed 

Google Scholar 

Kloe GE, Bailey D, Leurs R, de Esch IJP (2009) Transforming fragments into candidates: small becomes big in medicinal chemistry. Drug Discov Today 14:630–646. https://doi.org/10.1016/j.drudis.2009.03.009

Article 

PubMed 

Google Scholar 

Krimm I (2017) Applications of ligand and protein-observed NMR in ligand discovery. In: Applied biophysics for drug discovery. John Wiley & Sons, Ltd, Chichester, pp 175–195Chapter 

Google Scholar 

Ladbury JE, Klebe G, Freire E (2010) Adding calorimetric data to decision making in lead discovery: a hot tip. Nat Rev Drug Discov 9:23–27. https://doi.org/10.1038/nrd3054

Article 

CAS 

PubMed 

Google Scholar 

Larsson EA, Jansson A, Ng FM et al (2013) Fragment-based ligand design of novel potent inhibitors of tankyrases. J Med Chem 56:4497–4508. https://doi.org/10.1021/jm400211f

Article 

CAS 

PubMed 

Google Scholar 

Linke P, Amaning K, Maschberger M et al (2016) An automated microscale thermophoresis screening approach for fragment-based lead discovery. J Biomol Screen 21:414–421. https://doi.org/10.1177/1087057115618347

Article 

CAS 

PubMed 

Google Scholar 

Lippolis R, Angelis MD (2016) Proteomics and human diseases. J Proteomics Bioinform 9:63–74. https://doi.org/10.4172/jpb.1000391

Article 

CAS 

Google Scholar 

Lo MC, Aulabaugh A, Jin G et al (2004) Evaluation of fluorescence-based thermal shift assays for hit identification in drug discovery. Anal Biochem 332:153–159. https://doi.org/10.1016/j.ab.2004.04.031

Article 

CAS 

PubMed 

Google Scholar 

Maji B, Gangopadhyay SA, Lee M et al (2019) A high-throughput platform to identify small-molecule inhibitors of CRISPR-Cas9. Cell 177:1067–1079.e19. https://doi.org/10.1016/j.cell.2019.04.009

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Mann S, Ploux O (2006) 7,8-Diaminoperlargonic acid aminotransferase from Mycobacterium tuberculosis, a potential therapeutic target: characterization and inhibition studies. FEBS J. https://doi.org/10.1111/j.1742-4658.2006.05479.x

Article 

CAS 

PubMed 

Google Scholar 

Marciano DP, Dharmarajan V, Griffin PR (2014) HDX-MS guided drug discovery: small molecules and biopharmaceuticals. Curr Opin Struct Biol 28:105–111. https://doi.org/10.1016/j.sbi.2014.08.007

Article 

CAS 

PubMed 

Google Scholar 

Mashalidis EH, Śledå P, Lang S, Abell C (2013) A three-stage biophysical screening cascade for fragment-based drug discovery. Nat Protoc 8:2309–2324. https://doi.org/10.1038/nprot.2013.130

Article 

CAS 

PubMed 

Google Scholar 

Matulis D, Kranz JK, Salemme FR, Todd MJ (2005) Thermodynamic stability of carbonic anhydrase: measurements of binding affinity and stoichiometry using thermofluor. Biochemistry 44:5258–5266. https://doi.org/10.1021/bi048135v

Article 

CAS 

PubMed 

Google Scholar 

McClure SM, Ahl PL, Blue JT (2018) High throughput differential scanning fluorimetry (DSF) formulation screening with complementary dyes to assess protein unfolding and aggregation in presence of surfactants. Pharm Res 35:1–10. https://doi.org/10.1007/s11095-018-2361-1

Article 

CAS 

Google Scholar 

Menzen T, Friess W (2013) High-throughput melting-temperature analysis of a monoclonal antibody by differential scanning fluorimetry in the presence of surfactants. J Pharm Sci 102:415–428. https://doi.org/10.1002/jps.23405

Article 

CAS 

PubMed 

Google Scholar 

Mezzasalma TM, Kranz JK, Chan W et al (2007) Enhancing recombinant protein quality and yield by protein stability profiling. J Biomol Screen 12:418–428. https://doi.org/10.1177/1087057106297984

Article 

CAS 

PubMed 

Google Scholar 

Mizuno K, Boudko S, Engel J, Bachinger P (2010) Kinetic hysteresis in collagen folding. Biophys J 98:3004–3014. https://doi.org/10.1016/j.bpj.2010.03.019

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Molina DM, Jafari R, Ignatushchenko M et al (2013) Monitoring drug target engagement in cells and tissues using the cellular thermal shift assay. Science 341:84–87. https://doi.org/10.1126/science.1233606

Article 

CAS 

Google Scholar 

Moreau MJJ, Morin I, Schaeffer PM (2010) Quantitative determination of protein stability and ligand binding using a green fluorescent protein reporter system. Mol BioSyst 6:1285–1292. https://doi.org/10.1039/c002001j

Article 

CAS 

PubMed 

Google Scholar 

Murray CW, Rees DC (2009) The rise of fragment-based drug discovery. Nat Chem. https://doi.org/10.1038/nchem.217

Article 

CAS 

PubMed 

Google Scholar 

Navratilova I, Hopkins AL (2010) Fragment screening by surface plasmon resonance. ACS Med Chem Lett. https://doi.org/10.1021/ml900002k

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Neumann T, Junker H-D, Schmidt K, Sekul R (2007) SPR-based fragment screening: advantages and applications. Curr Top Med Chem 7:1630–1642. https://doi.org/10.2174/156802607782341073

Article 

CAS 

PubMed 

Google Scholar 

Newman J (2004) Novel buffer systems for macromolecular crystallization. Acta Crystallogr Sect D Biol Crystallogr 60:610–612. https://doi.org/10.1107/S0907444903029640

Article 

CAS 

Google Scholar 

Nielsen L, Khurana R, Coats A et al (2001) Effect of environmental factors on the kinetics of insulin fibril formation: elucidation of the molecular mechanism. Biochemistry 40:6036–6046. https://doi.org/10.1021/bi002555c

Article 

CAS 

PubMed 

Google Scholar 

Niesen FH, Berglund H, Vedadi M (2007) The use of differential scanning fluorimetry to detect ligand interactions that promote protein stability. Nat Protoc. https://doi.org/10.1038/nprot.2007.321

Article 

CAS 

PubMed 

Google Scholar 

Norin M, Sundström M (2002) Structural proteomics: developments in structure-to-function predictions. Trends BiotechnolOhlson S, Duong-Thi M-D (2018) Fragment screening for drug leads by weak affinity chromatography (WAC-MS). Methods 146:26–38. https://doi.org/10.1016/j.ymeth.2018.01.011

Article 

CAS 

PubMed 

Google Scholar 

Pantoliano MW, Whitlow M, Wood JF et al (1989) Large increases in general stability for subtilisin BPN’ through incremental changes in the free energy of unfolding. Biochemistry. https://doi.org/10.1021/bi00444a012

Article 

CAS 

PubMed 

Google Scholar 

Pantoliano MW, Horlick RA, Springer BA et al (1994) Multivalent ligand-receptor binding interactions in the fibroblast growth factor system produce a cooperative growth factor and heparin mechanism for receptor dimerization. Biochemistry. https://doi.org/10.1021/bi00200a003

Article 

CAS 

PubMed 

Google Scholar 

Pantoliano MW, Bone RF, Rhind AW, Salemme FR (1997) Microplate thermal shift assay apparatus for ligand development and multi-variable protein chemistry optimization. US Patent No. 6036920, 1997Pantoliano MW, Petrella EC, Kwasnoski JD et al (2001) High-density miniaturized thermal shift assays as a general strategy for drug discovery. J Biomol Screen 6:429–440. https://doi.org/10.1089/108705701753364922

Article 

CAS 

PubMed 

Google Scholar 

Park SW, Casalena D, Wilson D et al (2015) Target-based identification of whole-cell active inhibitors of biotin biosynthesis in Mycobacterium tuberculosis. Chem Biol 22:76–86. https://doi.org/10.1016/j.chem.biol.2014.xx.xxx

Article 

CAS 

PubMed 

Google Scholar 

Patel D, Bauman JD, Arnold E (2014) Advantages of crystallographic fragment screening: functional and mechanistic insights from a powerful platform for efficient drug discovery. Prog Biophys Mol Biol 116:92–100. https://doi.org/10.1016/j.pbiomolbio.2014.08.004

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Patton WF, Dai L, Ludlam A et al (2013) Dyes for analysis of protein aggregation. US Patent Application 20160280921, 2013Pedro L, Quinn R (2016) Native mass spectrometry in fragment-based drug discovery. Molecules 21:984. https://doi.org/10.3390/molecules21080984

Article 

CAS 

PubMed Central 

Google Scholar 

Qin S, Ren Y, Fu X et al (2015) Multiple ligand detection and affinity measurement by ultrafiltration and mass spectrometry analysis applied to fragment mixture screening. Anal Chim Acta 886:98–106. https://doi.org/10.1016/j.aca.2015.06.017

Article 

CAS 

PubMed 

Google Scholar 

Rainard JM, Pandarakalam GC, McElroy SP (2018) Using microscale thermophoresis to characterize hits from high-throughput screening: a European lead factory perspective. SLAS Discov Adv life Sci R D 23:225–241. https://doi.org/10.1177/2472555217744728

Article 

CAS 

Google Scholar 

Reinhard L, Mayerhofer H, Geerlof A et al (2013) Optimization of protein buffer cocktails using Thermofluor. Acta Crystallogr Sect F Struct Biol Cryst Commun 69:209–214. https://doi.org/10.1107/S1744309112051858

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Ren C, Bailey AO, VanderPorten E et al (2019) Quantitative determination of protein–ligand affinity by size exclusion chromatography directly coupled to high-resolution native mass spectrometry. Anal Chem 91:903–911. https://doi.org/10.1021/acs.analchem.8b03829

Article 

CAS 

PubMed 

Google Scholar 

Renaud JP, Chung CW, Danielson UH et al (2016) Biophysics in drug discovery: impact, challenges and opportunities. Nat Rev Drug Discov 15:679–698. https://doi.org/10.1038/nrd.2016.123

Article 

CAS 

PubMed 

Google Scholar 

Rosano GL, Ceccarelli EA (2014) Recombinant protein expression in Escherichia coli: advances and challenges. Front Microbiol 5:1–17. https://doi.org/10.3389/fmicb.2014.00172

Article 

Google Scholar 

Rumble C, Rich K, He G, Maroncelli M (2012) CCVJ is not a simple rotor probe. J Phys Chem A 116:10786–10792. https://doi.org/10.1021/jp309019g

Article 

CAS 

PubMed 

Google Scholar 

Sambandamurthy VK, Wang X, Chen B et al (2002) A pantothenate auxotroph of Mycobacterium tuberculosis is highly attenuated and protects mice against tuberculosis. Nat Med. https://doi.org/10.1038/nm765

Article 

CAS 

PubMed 

Google Scholar 

Sarciaux JM, Mansour S, Hageman MJ, Nail SL (1999) Effects of buffer composition and processing conditions on aggregation of bovine IgG during freeze-drying. J Pharm Sci 88:1354–1361. https://doi.org/10.1021/js980383n

Article 

CAS 

PubMed 

Google Scholar 

Sassetti CM, Rubin EJ (2003) Genetic requirements for mycobacterial survival during infection. Proc Natl Acad Sci U S A. https://doi.org/10.1073/pnas.2134250100

Article 

CAS 

Google Scholar 

Schmidt C (2010) GSK/Sirtris compounds dogged by assay artifacts. Nat Biotechnol 28:185–186. https://doi.org/10.1038/nbt0310-185

Article 

CAS 

PubMed 

Google Scholar 

Schumacher MA, Chinnam N, Ohashi T et al (2013) The structure of Irisin reveals a novel intersubunit β-sheet fibronectin type III (FNIII) dimer: implications for receptor activation. J Biol Chem 288:33738–33744. https://doi.org/10.1074/jbc.M113.516641

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Scott DE, Spry C, Abell C (2016) Differential scanning fluorimetry as part of a biophysical screening cascade. In: Erlanson DA, Jahnke W (eds) Fragment-based drug discovery lessons and outlook. Wiley-VCH Verlag, Weinheim, pp 139–172. https://doi.org/10.1002/9783527683604.ch07

Chapter 

Google Scholar 

Semisotnov GV, Rodionova NA, Razgulyaev OI et al (1991) Study of the “molten globule” intermediate state in protein folding by a hydrophobic fluorescent probe. Biopolymers 31:119–128. https://doi.org/10.1002/bip.360310111

Article 

CAS 

PubMed 

Google Scholar 

Shrake A, Ross PD (1992) Origins and consequences of ligand-induced multiphasic thermal protein denaturation. Biopolymers 32:925–940. https://doi.org/10.1002/bip.360320804

Article 

CAS 

PubMed 

Google Scholar 

Sørensen HP, Mortensen KK (2005) Advanced genetic strategies for recombinant protein expression in Escherichia coli. J Biotechnol 115:113–128. https://doi.org/10.1016/j.jbiotec.2004.08.004

Article 

CAS 

PubMed 

Google Scholar 

Szilágyi A, Závodszky P (2000) Structural differences between mesophilic, moderately thermophilic and extremely thermophilic protein subunits: results of a comprehensive survey. Structure. https://doi.org/10.1016/S0969-2126(00)00133-7

Article 

PubMed 

Google Scholar 

Valenti D, Neves JF, Cantrelle FX et al (2019) Set-up and screening of a fragment library targeting the 14-3-3 protein interface. Medchemcomm 10:1796–1802. https://doi.org/10.1039/c9md00215d

Article 

CAS 

PubMed 

Google Scholar 

Vedadi M, Niesen FH, Allali-Hassani A et al (2006) Chemical screening methods to identify ligands that promote protein stability, protein crystallization, and structure determination. Proc Natl Acad Sci 103:15835–15840. https://doi.org/10.1073/pnas.0605224103

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Wakayama R, Uchiyama S, Hall D (2019) Ionic liquids and protein folding-old tricks for new solvents. Biophys Rev 11:209–225. https://doi.org/10.1007/s12551-019-00509-2

Article 

PubMed 

PubMed Central 

Google Scholar 

Wang Y, Van Oosterwijk N, Ali AM et al (2017) A systematic protein refolding screen method using the DGR approach reveals that time and secondary TSA are essential variables. Sci Rep 7:1–10. https://doi.org/10.1038/s41598-017-09687-z

Article 

CAS 

PubMed 

PubMed Central 

Google Scholar 

Wang S, Dong G, Sheng C (2019) Structural simplification: an efficient strategy in lead optimization. Acta Pharm Sin B 9:880–901. https://doi.org/10.1016/j.apsb.2019.05.004

Article 

PubMed 

PubMed Central 

Google Scholar 

Wartchow CA, Podlaski F, Li S et al (2011) Biosensor-based small molecule fragment screening with biolayer interferometry. J Comput Aided Mol Des 25:669–676. https://doi.org/10.1007/s10822-011-9439-8

Article 

CAS 

PubMed 

Google Scholar 

Weber PC, Pantoliano MW, Simons DM, Salemme FR (1994) Structure-based design of synthetic azobenzene ligands for streptavidin. J Am Chem Soc 116:2717–2724. https://doi.org/10.1021/ja00086a004

Article 

CAS 

Google Scholar 

Wienken CJ, Baaske P, Rothbauer U et al (2010) Protein-binding assays in biological liquids using microscale thermophoresis. Nat Commun. https://doi.org/10.1038/ncomms1093

Wingfield PT (2015) Overview of the purification of recombinant proteins. Curr Protoc Protein Sci 80:6.1.1–6.1.35. https://doi.org/10.1002/0471140864.ps0601s80

World Health Organization (2018) Global tuberculosis report 2018 - Geneva. Retrieved from: https://apps.who.int/iris/bitstream/handle/10665/274453/9789241565646-eng.pdf. Accessed 5 Dec 2019Xu M, Liu C, Zhou M et al (2016) Screening of small-molecule inhibitors of protein–protein interaction with capillary electrophoresis frontal analysis. Anal Chem 88:8050–8057. https://doi.org/10.1021/acs.analchem.6b01430

Article 

CAS 

PubMed 

Google Scholar 

Xu S, Aguilar A, Xu T et al (2018) Design of the first-in-class, highly potent irreversible inhibitor targeting the menin-MLL protein–protein interaction. Angew Chemie - Int Ed. https://doi.org/10.1002/anie.201711828

Article 

CAS 

PubMed 

Google Scholar 

Download referencesAcknowledgmentsWe would like to thank the kind support and access to sample application from NanoTemper (München, Germany).Author informationAuthors and AffiliationsStructure Biology in Drug Design, Drug Design Group XB20, Departments of Pharmacy, University of Groningen, Groningen, The NetherlandsKai Gao, Rick Oerlemans & Matthew R. GrovesAuthorsKai GaoView author publicationsYou can also search for this author in

PubMed Google ScholarRick OerlemansView author publicationsYou can also search for this author in

PubMed Google ScholarMatthew R. GrovesView author publicationsYou can also search for this author in

PubMed Google ScholarCorresponding authorCorrespondence to

Matthew R. Groves.Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissionsAbout this articleCite this articleGao, K., Oerlemans, R. & Groves, M.R. Theory and applications of differential scanning fluorimetry in early-stage drug discovery.

Biophys Rev 12, 85–104 (2020). https://doi.org/10.1007/s12551-020-00619-2Download citationReceived: 04 December 2019Accepted: 08 January 2020Published: 31 January 2020Issue Date: February 2020DOI: https://doi.org/10.1007/s12551-020-00619-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy to clipboard

Provided by the Springer Nature SharedIt content-sharing initiative

KeywordsThermal stabilityFoldingUnfoldingRefoldingFluorimetryLigands screeningCrystallizationBuffer optimization

Use our pre-submission checklist

Avoid common mistakes on your manuscript.

Advertisement

Search

Search by keyword or author

Search

Navigation

Find a journal

Publish with us

Track your research

Discover content

Journals A-Z

Books A-Z

Publish with us

Publish your research

Open access publishing

Products and services

Our products

Librarians

Societies

Partners and advertisers

Our imprints

Springer

Nature Portfolio

BMC

Palgrave Macmillan

Apress

Your privacy choices/Manage cookies

Your US state privacy rights

Accessibility statement

Terms and conditions

Privacy policy

Help and support

49.157.13.121

Not affiliated

© 2024 Springer Nature